Next Article in Journal
Chatter Stability Prediction and Process Parameters’ Optimization of Milling Considering Uncertain Tool Information
Next Article in Special Issue
Base Catalysis of Sodium Salts of [Ta6−xNbxO19]8− Mixed-Oxide Clusters
Previous Article in Journal
Study on the Face Stability of a Metro Tunnel in a Silty Clay Layer Constructed Using the Full-Face Method
Previous Article in Special Issue
Relativistic Corrections to Phase Shift Calculation in the GNXAS Package
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Extracting Local Symmetry of Mono-Atomic Systems from Extended X-ray Absorption Fine Structure Using Deep Neural Networks

1
Aichi Synchrotron Radiation Center, Aichi, Seto 489-0965, Japan
2
Kyushu Synchrotron Light Research Center, Saga, Tosu 841-0005, Japan
*
Author to whom correspondence should be addressed.
Symmetry 2021, 13(6), 1070; https://doi.org/10.3390/sym13061070
Submission received: 29 May 2021 / Revised: 6 June 2021 / Accepted: 11 June 2021 / Published: 15 June 2021
(This article belongs to the Special Issue X-ray Absorption Fine Structure and Symmetry)

Abstract

:
In recent years, neural networks have become a new method for the analysis of extended X-ray absorption fine structure data. Due to its sensitivity to local structure, X-ray absorption spectroscopy is often used to study disordered systems and one of its more interesting property is the sensitivity not only to pair distribution function, but also to three-body distribution, which contains information on the local symmetry. In this study, by considering the case of Ni, we show that by using neural networks, it is possible to obtain not only the radial distribution function, but also the bond angle distribution between the first nearest-neighbors. Additionally, by adding appropriate configurations in the dataset used for training, we show that the neural network is able to analyze also data from disordered phases (liquid and undercooled state), detecting small changes in the local ordering compatible with results obtained through other methods.

1. Introduction

Machine Learning (ML) methods are becoming widely used to tackle different scientific problems [1,2]. This popularity comes not only thanks to the advancement in the field and the success of the results, but also due to the availability of various open-source packages, which make them accessible even to non-expert users. In recent years, various studies have been conducted to the use of those methods connected to X-ray absorption spectroscopy (XAS) [3,4,5,6,7,8,9,10,11,12,13,14,15]. XAS is an experimental technique where, by measuring the absorbance of a sample as a function of energy around core-electrons edges of elements present in the sample, we can obtain information about the local structure [16,17]. The spectra is commonly divided into two regions, due to both experimental and theoretical reasons, depending on the energy range we consider: X-ray Absorption Near Edge Structure (XANES) for the near-edge region (up to about 50 eV after the edge) and Extended X-ray Absorption Fine Structure (EXAFS) for the region above 50 eV and beyond. One the main advantages of XAS is that it does not require sample crystallinity, meaning it can be used to study a vast range of materials, such as molecules [18,19,20], amorphous systems [21,22,23,24] and nano-materials [25,26,27,28].
In this paper, we focus our attention on EXAFS. The EXAFS signal can be described in terms of the n-body distribution functions g n [29,30]:
χ ( k ) = 0 d r 4 π r 2 ρ g 2 ( r ) γ ( 2 ) ( r , k ) + d r 1 d r 2 d ϕ 8 π 2 r 1 2 r 2 2 sin ( ϕ ) ρ 2 g 3 ( r 1 , r 2 , ϕ ) γ ( 3 ) ( r 1 , r 2 , ϕ , k ) + O ( g 4 ) ,
where k is the photoelectron wave-vector k = 2 m ( E E e ) / defined from the edge’s energy and γ ( n ) are the irreducible n-body contributions to EXAFS, containing the MS terms associated with a given geometry. The n-body distributions are statistical quantities proportional to the probability of finding atoms in a given configuration and are used to describe the structure of the system [31]. Due to mean-free-path effects of the excited photo-electron, two-body and three-body terms are sufficient to describe experimental EXAFS signals and contributions from higher order distributions can be ignored. Direct extraction of the structural functions from Equation (1) is a known ill-posed problems and various strategies have been used over the years to solve this problem. The standard method of analysis, so called ‘peak-fitting’ approach, consists in describing the distribution functions as a sum of peaks and then refine the peak parameters by a least-square fitting with the experimental data. This method has the drawback of having to know a priori the peak distributions and additional constraints are required to treat disordered systems [32,33]. Other methods of analysis based on regularization techniques [34,35] have also been used, although limited to single scattering terms. Reverse Monte Carlo (RMC) [36] is a structural modeling method that can also be used to analyze EXAFS data [37,38,39], but being a simulation method, it may be costly in terms of time and also the density of the system required. On the other hand, it permits us to calculate, from the obtained coordinates, radial distributions functions, and bond-angle distributions and look for specific geometries [40,41].
As mentioned before, ML, more precisely Neural Networks (NN), has recently came out as another method to analyze EXAFS data [13,42]. A NN consists of many “neurons” stacked into layers, with the output of one layer acting as the input for the next. A neuron performs a linear transformation of the inputs with given weights and pass it to a step-wise activation function, which gives the output. The weights used are adjusted during the training of the NN, performed on a dataset for the problem that we want to study. For EXAFS, computer simulations, like Molecular Dynamics (MD) [43] are used to create the dataset, since it is not feasible using only experimental data. Previous works mainly focused on crystalline configurations, we want to investigate whether the same method could be used to also investigate disordered phases and whether we could obtain, other than the radial distribution function, also the bond-angle distributions, which contain information on the local geometries. In this work, we consider the case of metallic Ni, for which also the liquid phase has been previously measured [44].

2. Methods

2.1. MD Simulations

To create the dataset for the training of the NN, we run several MD simulations using the LAMMPS code [45,46]. Four different crystal configurations were taken into account: face-centered cubic (fcc), hexagonal close packed (hcp), body-centered cubic (bcc) and face-centered diamond-cubic (dia). Simulations were done in an NVT ensemble using different modified embedded-atom potentials ([47] for fcc, hcp and bcc, [48] for dia) depending on the starting crystal configuration to avoid transition to a different structure. Since MD configurations are used only to establish the relation between the structure and the EXAFS signal, it is not necessary in this case that they reproduce real nickel configurations. For all structures, the lattice parameter was chosen to have approximately the first nearest-neighbors at the distance of room temperature nickel ∼2.49 Å [49]. Simulations for each structure were run from 60 to 1500 K with a step Δ T = 20 K, for each temperature the simulation run for 50 ps to ensure equilibration. After that, the atomic configuration was saved every 1 ps for a total of 21 times. In order to save computational time, these configurations were rescaled to change the first nearest-neighbors distance with a step Δ r 0.015 Å from −10 to +10 times, effectively obtaining 21 different volume configurations for a given temperature.
In addition to crystal configurations, disordered liquid configurations were created. From an initial fcc configuration with a lattice parameter chosen to reproduce the liquid density [50], NVT simulations were performed at 2000 K for 500 ps, ensuring that the final configuration consisted in a melted system with no memory of the initial crystal structure. This final configuration was then used as a starting point for several simulation at different temperatures from 2000 to 1500 K with a step Δ T = 20 K that were run 100 ps to ensure equilibration. Similar to the crystal case, volume variations were done by rescaling different configurations for each temperature.
Number of atoms were chosen to have sufficient statistics, but limited to reduce computational time. In the crystal cases, we used 864 atoms for fcc and hcp (6 × 6 × 6 unit cells), 1024 atoms for bcc (8 × 8 × 8 unit cells) and 1000 atoms for dia (5 × 5 × 5 unit cells). For the liquid configurations, 4000 atoms were used instead: due to the disorder, more atoms were required to have sufficient statistics. In total, we obtained more than 6000 configurations, from which we calculated the radial distribution function (RDF) n ( r ) = 4 π r 2 ρ g ( r ) (cfr. Equtation (1)) using a binning of 0.05 Å, the bond angle distribution (BAD), using the minimum after the first peak as a cut-off (for bcc, the first two peaks were included, since they merge at high temperature), and the EXAFS signal.
While for crystals the nearest neighbor peak is clearly defined and so is the bond angle distribution, the definition for the liquid case is not unique [51]. The most common method is the use of a distance cut-off positioned around the first minimum, but small variations of this value can slightly change the distribution, due to more or less atoms being included in the calculations, which changes the height of the distribution. For comparison then, the normalized distribution (the integral is normalized to a constant) is used instead.

2.2. EXAFS Calculation

The χ ( k ) EXAFS signal for a configuration is calculated as the average of the signal coming from all atoms (ensemble average) using the suite of programs GNXAS [52,53]:
χ ( k ) = i γ ( 2 ) ( 0 , i ) + ( i , j ) γ ( 3 ) ( 0 , i , j ) over   all   atoms ,
where γ ( 2 ) and γ ( 3 ) are the irreducible 2-body and 3-body signals associated with the photo-absorbing atom 0 and its surrounding atoms up to a path cut-off. These signals are calculated by continued fraction expansion [54], containing all the significant multiple-scattering terms associated with that geometry. The use of a path cut-off is justified by the mean-free-path effects of the excited photo-electron, which makes the signal associated to distant coordination shells negligible. The cut-off used for crystal configuration was 8 Å and decreased to 6 Å for liquid configurations, due to the increased disorder. In both cases, it was confirmed that the use of a higher cut-off did not change the calculated EXAFS signal.
The energy shift for the theoretical calculations was chosen randomly between −5 to +10 eV for each configuration. It has been shown [8] that using this strategy makes the NN trained on the dataset insensitive to this parameter.

2.3. NN Architecture and Training

For the NN, a short Python3 [55] program was written, making use of the TensorFlow [56] and Keras [57] packages. The architecture consisted, other than the input layer ( k 2 -weighted EXAFS signal) and output layers ( n ( r ) and BAD), in 4 hidden dense layers of 1000 neurons each with a rectified linear activation function (ReLU). Training was performed using the Adam optimizer [58] with a learning rate of 10 4 and the metrics used were the mean squared errors for both quantities, scaled to have roughly the same weight. Early stopping with a patience of 50 steps was also used. All these parameters were obtained by grid search and through trial-and-error. For all optimization procedures, the dataset was divided in three chunks chosen randomly, following common practices of machine learning: a train set, a validation set and a test set, with a ratio of roughly 3:1:1, respectively. The train set was used to for the training of the weights in the NN, the validation set to check the training procedure and enforce the early stopping, and finally the test set to evaluate the performance of the NN. Parameters that obtained lowest loss on the test set were chosen. The use of Dropout [59] was also tested, but was found to not increase performance in our case.
For the training of the NNs used for data analysis, again 20% of the dataset was used for the test split, and the rest was divided into 5 equal parts. In a procedure similar to cross validation [60], one of the parts was used as validation set, while the others were used for training in rotation, for a total of 5 trained NNs. Results from each NN are then averaged and the standard deviation used to estimate errors.

3. Results

3.1. Test Dataset

We first checked the performance on the test dataset, which has not been used for training. In Figure 1, we show in the top left the mean square error (MSE), ordered in descending order, for the worst 200 cases, and some selected configurations, one for each of the different type of structure we included in the dataset. For each we compare the RDF and BAD predicted by the NN from the calculated EXAFS signal with the ones from the model. Even in the cases where the MSE is high, like (a) and (b), the overall features of the distributions are correctly predicted and additional peaks, if present, have standard deviations that exceed their value. All the different type of structure are correctly identified.

3.2. Experimental Data

We then tested the NN on experimental data for Ni. XAFS spectra from a Ni thin-film was collected in transmission geometry at the beamline BL11 [61] of the Saga Light Source. The EXAFS signal (Figure 2) was extracted using the program jesf from the GNXAS package. With respect to the simulated data used for training, experimental data contain noise produced by the ionization chamber detectors, but we have found that the noise does not much affect the results. We tested this by manually increasing the noise present in the data: while the standard deviation increased as the noise increased, the results remained consistent until the signal-to-noise ratio approached 1.
The data was fed to the NN, and the result obtained for the RDF is shown in Figure 3, compared with 4000 atoms NVT MD simulation at 300 K using the same potential used for constructing the dataset [47]. Peak position from X-ray diffraction (XRD) [49] and RDF obtained by RMC [42] are shown for comparison. We can see that the MD simulation in this case underestimates the thermal vibrations, resulting in slightly narrower and higher peaks with respect to the RDF obtained by NN, which instead agrees very well with previously published data [42]. Other than RDF, we obtained the BAD shown in Figure 4.
As with the RDF, also here we can see that the thermal disorder for the MD is lower respect to the prediction of the NN, but the angle’s positions are correctly estimated. Doing the integration of each peak gives the number of atoms with that specific geometry: for the NN result, these values are shown in Figure 4 together with estimated error bars from the standard deviation. The correct coordinations for a fcc structure (shown also as gray bars) are obtained.
We then used the NN to analyze data of undercooled and liquid Ni previously published [44]. In Figure 5, we report the obtained RDF obtained by NN, compared with NVT MD simulations at experimental density [50] and RMC analysis of the same data.
Some differences are present: most noticeable the right side of the first peak is broader for RMC. Since it is a method based on random movements of the atoms, it is known to give more disordered configurations that could result, like in this case, to broader peaks. Secondly, the first peak from the NN is slightly shifted to lower distance in both cases, more visible in the liquid data, respect to MD and RMC results. This difference could be explained by the correlation between the average distance and the energy shift parameter Δ E . As explained in Section 2.2, this parameter was randomly shifted in the dataset, so that the NN became insensitive to it: while this strategy works well for the crystal phase thanks to various multiple-scattering contribution to the EXAFS signal, for the liquid phase, where most terms become zero, it may not be sufficient. The shift is, in any case, of the order of 0.02 Å. On the right side of Figure 5, the first peak obtained by the NN for the two cases are compared: clear changes are visible that predict the correct temperature behavior, even if the two experimental signals difference is quite small [44].
On Figure 6, we show the normalized BAD distribution from NN, again compared with MD and RMC. As anticipated in Section 2.1, in the liquid case, the nearest neighbors are not unequivocally defined and the use of a cut-off distance around the first minimum is usually used. Small variations of the cut-off, which increases/decreases atoms considered, result in magnitude variation of the BAD. Therefore, for comparing purposes, the use of the normalized distribution is better suited. We can see that there is a good agreement between NN and RMC results for the peak around 60 , while some differences are present in the 90–120 region. This is somewhat expected, since contribution to the EXAFS signal as a function of angle has minimum around this angle (see Appendix A, Figure A1) and so the prediction given by the NN is biased by the dataset used: we can see that in this region indeed the distributions obtained from the NN resemble the ones from MD simulations. Nevertheless, the agreement is overall good and the NN is able to predict the small differences between the two signals, which is an indication of changes in the local geometry of the two phases.

4. Discussion and Conclusions

We have shown that the NN method can be used to obtain not just radial distribution function, but also bond angle distribution, which gives information on the local symmetry of nearest neighboring atoms. Additionally, the method can be also used to study non-crystalline configurations, like liquids and amorphous materials. While this method is certainly powerful and permits quick and detailed analysis also by non-experts, results should be compared to additional methods or experimental data for validation and exclude possible biases from the dataset. To improve upon this point, datasets could be expanded, including additional types of structures, and shared, since their creation is the biggest bottleneck in terms of time. Extension to multi-component systems is technically possible, but further investigations in this direction are required for disordered systems.

Author Contributions

Conceptualization, F.I. and T.O.; methodology, F.I.; software, F.I.; validation, H.S. and T.O.; formal analysis, F.I.; investigation, H.S.; resources, H.S. and T.O.; data curation, F.I.; writing—original draft preparation, F.I.; writing—review and editing, H.S. and T.O.; visualization, F.I., H.S. and T.O; supervision, F.I.; project administration, T.O.; funding acquisition, T.O. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Japan Science and Technology Agency Core Research for Evolutional Science and Technology program grant number JPMJCR1861.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available from the corresponding author upon reasonable request.

Acknowledgments

We thank Andrea Di Cicco (University of Camerino) for permission to use data from Reference [44]. Part of the computation was carried out using the computer resources offered by the Research Institute for Information Technology, Kyushu University.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
MLMachine Learning
XASX-ray absorption spectroscopy
XANESX-ray absorption near edge structure
EXAFSExtended x-ray absorption fine structure
NNNeural Network
MDMolecular Dynamics
RDFRadial distribution function
BADBond-angle distribution

Appendix A

Amplitude of the γ ( 3 ) signals as a function of the angle between two Ni atoms at distance 2.487 Å from the photo-absorbing atom. The increase at 180 is known as the focusing effect given by the co-linear configuration [29,62].
Figure A1. Amplitude of γ ( 3 ) as a function of the angle.
Figure A1. Amplitude of γ ( 3 ) as a function of the angle.
Symmetry 13 01070 g0a1

References

  1. Mehta, P.; Bukov, M.; Wang, C.H.; Day, A.G.; Richardson, C.; Fisher, C.K.; Schwab, D.J. A high-bias, low-variance introduction to Machine Learning for physicists. Phys. Rep. 2019, 810, 1–124. [Google Scholar] [CrossRef] [PubMed]
  2. Carleo, G.; Cirac, I.; Cranmer, K.; Daudet, L.; Schuld, M.; Tishby, N.; Vogt-Maranto, L.; Zdeborová, L. Machine learning and the physical sciences. Rev. Mod. Phys. 2019, 91, 45002. [Google Scholar] [CrossRef] [Green Version]
  3. Timoshenko, J.; Lu, D.; Lin, Y.; Frenkel, A.I. Supervised Machine-Learning-Based Determination of Three-Dimensional Structure of Metallic Nanoparticles. J. Phys. Chem. Lett. 2017, 8, 5091–5098. [Google Scholar] [CrossRef]
  4. Guda, A.A.; Guda, S.A.; Lomachenko, K.A.; Soldatov, M.A.; Pankin, I.A.; Soldatov, A.V.; Braglia, L.; Bugaev, A.L.; Martini, A.; Signorile, M.; et al. Quantitative structural determination of active sites from in situ and operando XANES spectra: From standard ab initio simulations to chemometric and machine learning approaches. Catal. Today 2019, 336, 3–21. [Google Scholar] [CrossRef]
  5. Timoshenko, J.; Wrasman, C.J.; Luneau, M.; Shirman, T.; Cargnello, M.; Bare, S.R.; Aizenberg, J.; Friend, C.M.; Frenkel, A.I. Probing Atomic Distributions in Mono- and Bimetallic Nanoparticles by Supervised Machine Learning. Nano Lett. 2018, 19, 520–529. [Google Scholar] [CrossRef]
  6. Zheng, C.; Mathew, K.; Chen, C.; Chen, Y.; Tang, H.; Dozier, A.; Kas, J.J.; Vila, F.D.; Rehr, J.J.; Piper, L.F.; et al. Automated generation and ensemble-learned matching of X-ray absorption spectra. NPJ Comput. Mater. 2018, 4, 1–9. [Google Scholar] [CrossRef] [Green Version]
  7. Suzuki, Y.; Hino, H.; Kotsugi, M.; Ono, K. Automated estimation of materials parameter from X-ray absorption and electron energy-loss spectra with similarity measures. NPJ Comput. Mater. 2019, 5, 1–7. [Google Scholar] [CrossRef]
  8. Timoshenko, J.; Ahmadi, M.; Roldan Cuenya, B. Is There a Negative Thermal Expansion in Supported Metal Nanoparticles? An in Situ X-ray Absorption Study Coupled with Neural Network Analysis. J. Phys. Chem. C 2019, 123, 20549–20604. [Google Scholar] [CrossRef] [Green Version]
  9. Kiyohara, S.; Mizoguchi, T. Radial Distribution Function from X-ray Absorption near Edge Structure with an Artificial Neural Network. J. Phys. Soc. Jpn. 2020, 89, 1–4. [Google Scholar] [CrossRef]
  10. Marcella, N.; Liu, Y.; Timoshenko, J.; Guan, E.; Luneau, M.; Shirman, T.; Plonka, A.M.; Van Der Hoeven, J.E.; Aizenberg, J.; Friend, C.M.; et al. Neural network assisted analysis of bimetallic nanocatalysts using X-ray absorption near edge structure spectroscopy. Phys. Chem. Chem. Phys. 2020, 22, 18902–18910. [Google Scholar] [CrossRef]
  11. Carbone, M.R.; Topsakal, M.; Lu, D.; Yoo, S. Machine-Learning X-ray Absorption Spectra to Quantitative Accuracy. Phys. Rev. Lett. 2020, 124, 156401. [Google Scholar] [CrossRef] [Green Version]
  12. Iwamitsu, K.; Yokota, T.; Murata, K.; Kamezaki, M.; Mizumaki, M.; Uruga, T.; Akai, I. Spectral Analysis of X-ray Absorption Near Edge Structure in α-Fe2O3 Based on Bayesian Spectroscopy. Phys. Status Solidi (B) Basic Res. 2020, 257, 2–7. [Google Scholar] [CrossRef]
  13. Timoshenko, J.; Jeon, H.S.; Sinev, I.; Haase, F.T.; Herzog, A.; Roldan Cuenya, B. Linking the evolution of catalytic properties and structural changes in copper-zinc nanocatalysts using: Operando EXAFS and neural-networks. Chem. Sci. 2020, 11, 3727–3736. [Google Scholar] [CrossRef] [Green Version]
  14. Terry, J.; Lau, M.L.; Sun, J.; Xu, C.; Hendricks, B.; Kise, J.; Lnu, M.; Bagade, S.; Shah, S.; Makhijani, P.; et al. Analysis of extended X-ray absorption fine structure (EXAFS) data using artificial intelligence techniques. Appl. Surf. Sci. 2021, 547, 149059. [Google Scholar] [CrossRef]
  15. Zheng, C.; Chen, C.; Chen, Y.; Ong, S.P. Random Forest Models for Accurate Identification of Coordination Environments from X-ray Absorption Near-Edge Structure. Patterns 2020, 1, 100013. [Google Scholar] [CrossRef] [PubMed]
  16. Bunker, G. Introduction to XAFS; Cambridge University Press: Cambridge, UK, 2010. [Google Scholar] [CrossRef]
  17. Calvin, S. XAFS for Everyone; CRC Press: Boca Raton, FL, USA, 2013. [Google Scholar] [CrossRef]
  18. Filipponi, A.; D’ Angelo, P. Accurate determination of molecular structures by X-ray absorption spectroscopy. J. Chem. Phys. 1998, 109, 5356. [Google Scholar] [CrossRef]
  19. Di Cicco, A.; Iesari, F.; Trapananti, A.; D’Angelo, P.; Filipponi, A. Structure and atomic correlations in molecular systems probed by XAS reverse Monte Carlo refinement. J. Chem. Phys. 2018, 148, 094307. [Google Scholar] [CrossRef] [Green Version]
  20. Lindblad, R.; Kjellsson, L.; Couto, R.C.; Timm, M.; Bülow, C.; Zamudio-Bayer, V.; Lundberg, M.; Von Issendorff, B.; Lau, J.T.; Sorensen, S.L.; et al. X-ray Absorption Spectrum of the N2+ Molecular Ion. Phys. Rev. Lett. 2020, 124, 203001. [Google Scholar] [CrossRef]
  21. Filipponi, A.; Di Cicco, A. Short-range order in crystalline, amorphous, liquid, and supercooled germanium probed by X-ray absorption spectroscopy. Phys. Rev. B 1995, 51, 12322–12336. [Google Scholar] [CrossRef]
  22. Di Cicco, A.; Trapananti, A.; Principi, E.; De Panfilis, S.; Filipponi, A. Polymorphism and metastable phenomena in liquid tin under pressure. Appl. Phys. Lett. 2006, 89, 221912. [Google Scholar] [CrossRef]
  23. Coppari, F.; Chervin, J.; Congeduti, A.; Lazzeri, M.; Polian, A.; Principi, E.; Di Cicco, A. Pressure-induced phase transitions in amorphous and metastable crystalline germanium by Raman scattering, X-ray spectroscopy, and ab initio calculations. Phys. Rev. B 2009, 80, 115213. [Google Scholar] [CrossRef]
  24. Properzi, L.; Santoro, M.; Minicucci, M.; Iesari, F.; Ciambezi, M.; Nataf, L.; Le Godec, Y.; Irifune, T.; Baudelet, F.; Di Cicco, A. Structural evolution mechanisms of amorphous and liquid As2Se3 at high pressures. Phys. Rev. B Condens. Matter Mater. Phys. 2016, 93, 1–8. [Google Scholar] [CrossRef]
  25. Sun, Y.; Frenkel, A.I.; Isseroff, R.; Shonbrun, C.; Forman, M.; Shin, K.; Koga, T.; White, H.; Zhang, L.; Zhu, Y.; et al. Characterization of palladium nanoparticles by using X-ray reflectivity, EXAFS, and electron microscopy. Langmuir 2006, 22, 807–816. [Google Scholar] [CrossRef] [PubMed]
  26. Ikemoto, H.; Miyanaga, T. Extended X-ray absorption fine structure study of local structure and atomic correlations of tellurium nanoparticles. Phys. Rev. Lett. 2007, 99, 165503. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Kuzmin, A.; Chaboy, J. EXAFS and XANES analysis of oxides at the nanoscale. IUCrJ 2014, 1, 571–589. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Ikemoto, H.; Fujimori, T.; Miyanaga, T.; Kato, S.; Iesari, F.; Urita, K. Structures of Isolated Tellurium Chains Encapsulated Inside Carbon Nanotube. J. Phys. Chem. C 2020, 124, 26043–26047. [Google Scholar] [CrossRef]
  29. Filipponi, A.; Di Cicco, A.; Natoli, C.R. X-ray absorption spectroscopy and n-body distribution functions in condensed matter (I): Theory. Phys. Rev. B 1995, 52, 15122–15134. [Google Scholar] [CrossRef] [PubMed]
  30. Filipponi, A.; Di Cicco, A. X-ray absorption spectroscopy and n-body distribution functions in condensed matter (II): Data-analysis and applications. Phys. Rev. B 1995, 52, 15135. [Google Scholar] [CrossRef]
  31. Allen, M.P.; Tildesley, D.J. Computer Simulation of Liquids; Clarendon Press: Oxford, UK, 1987. [Google Scholar]
  32. Filipponi, A. EXAFS for liquids. J. Phys. Condens. Matter 2001, 13, R23. [Google Scholar] [CrossRef]
  33. Trapananti, A.; Di Cicco, A. Probing the local structure of liquid binary mixtures by X-ray absorption spectroscopy. Phys. Rev. B 2004, 70, 014101. [Google Scholar] [CrossRef]
  34. Babanov, Y.A.; Vasin, V.V.; Ageev, A.L.; Ershov, N.V. A New Interpretation of EXAFS Spectra in Real Space. I. General Formalism. Phys. Stat. Sol. 1981, 105, 747. [Google Scholar] [CrossRef]
  35. Akai, I.; Iwamitsu, K.; Igarashi, Y.; Okada, M.; Setoyama, H.; Okajima, T.; Hirai, Y. Sparse Modeling of an Extended X-ray Absorption Fine-Structure Spectrum Based on a Single-Scattering Formalism. J. Phys. Soc. Jpn. 2018, 87, 074003. [Google Scholar] [CrossRef]
  36. McGreevy, R.L.; Pusztai, L. Reverse Monte Carlo simulation: A new technique for the determination of disordered structures. Mol. Sim. 1988, 1, 359–367. [Google Scholar] [CrossRef]
  37. Gurman, S.J.; McGreevy, R.L. Reverse Monte Carlo simulation for the analysis of EXAFS data. J. Phys. Condens. Matter 1990, 2, 9463–9473. [Google Scholar] [CrossRef]
  38. Di Cicco, A.; Trapananti, A. Reverse Monte Carlo refinement of molecular and condensed systems by X-ray absorption spctroscopy. J. Phys. Condens. Matter 2005, 17, S135–S144. [Google Scholar] [CrossRef]
  39. Timoshenko, J.; Kuzmin, A.; Purans, J. An efficient implementation of the reverse Monte Carlo method for EXAFS analysis in crystalline materials. J. Phys. Conf. Ser. 2013, 430, 012012. [Google Scholar] [CrossRef]
  40. Di Cicco, A.; Trapananti, A.; Faggioni, S.; Filipponi, A. Is there icosahedral ordering in liquid and undercooled metals? Phys. Rev. Lett. 2003, 91, 135505–1–135505–4. [Google Scholar] [CrossRef]
  41. Iesari, F.; Trapananti, A.; De Panfilis, S.; Di Cicco, A. Structure of liquid In20Sn80 at high temperature: A XAS study. Radiat. Phys. Chem. 2020, 175, 108089. [Google Scholar] [CrossRef]
  42. Timoshenko, J.; Anspoks, A.; Cintins, A.; Kuzmin, A.; Purans, J.; Frenkel, A.I. Neural Network Approach for Characterizing Structural Transformations by X-ray Absorption Fine Structure Spectroscopy. Phys. Rev. Lett. 2018, 120, 225502. [Google Scholar] [CrossRef] [Green Version]
  43. Alder, B.J.; Wainwright, T.E. Studies in Molecular Dynamics. I. General Method. J. Chem. Phys. 1959, 31, 459–466. [Google Scholar] [CrossRef] [Green Version]
  44. Di Cicco, A.; Iesari, F.; De Panfilis, S.; Celino, M.; Giusepponi, S.; Filipponi, A. Local fivefold symmetry in liquid and undercooled Ni probed by X-ray absorption spectroscopy and computer simulations. Phys. Rev. B 2014, 89, 060102. [Google Scholar] [CrossRef]
  45. Plimpton, S. Fast Parallel Algorithms for Short-Range Molecular Dynamics. J. Comput. Phys. 1995, 117, 1–19. [Google Scholar] [CrossRef] [Green Version]
  46. LAMMPS Website. Available online: http://lammps.sandia.gov (accessed on 28 May 2021).
  47. Etesami, S.A.; Asadi, E. Molecular dynamics for near melting temperatures simulations of metals using modified embedded-atom method. J. Phys. Chem. Solids 2018, 112, 61–72. [Google Scholar] [CrossRef]
  48. Kim, E.H.; Shin, Y.H.; Lee, B.J. A modified embedded-atom method interatomic potential for Germanium. Calphad 2008, 32, 34–42. [Google Scholar] [CrossRef]
  49. Otte, H.M. Lattice Parameter Determinations with an X-ray Spectrogoniometer by the Debye-Scherrer Method and the Effect of Specimen Condition. J. Appl. Phys. 1961, 32, 1536–1546. [Google Scholar] [CrossRef]
  50. Brillo, J.; Egry, I. Density determination of liquid copper, nickel, and their alloys. Int. J. Thermophys. 2003, 24, 1155. [Google Scholar] [CrossRef]
  51. Waseda, Y. The Structure of Non-Crystalline Materials; McGraw-Hill: New York, NY, USA, 1980. [Google Scholar]
  52. Di Cicco, A. (Ed.) GNXAS. Extended Suite of Programs for Advanced X-ray Absorption Data-Analysis: Methodology and Practice; TASK Publishing: Gdansk, Poland, 2009. [Google Scholar]
  53. GNXAS Website. Available online: http://gnxas.unicam.it (accessed on 28 May 2021).
  54. Filipponi, A. Continued fraction expansion for the X-ray absorption cross section. J. Phys. Condens. Matter 1991, 3, 6489–6507. [Google Scholar] [CrossRef]
  55. Van Rossum, G.; de Boer, J. Interactively Testing Remote Servers Using the Python Programming Language. CWI Q. 1991, 4, 283–303. [Google Scholar]
  56. Abadi, M.; Agarwal, A.; Barham, P.; Brevdo, E.; Chen, Z.; Citro, C.; Corrado, G.S.; Davis, A.; Dean, J.; Devin, M.; et al. TensorFlow: Large-Scale Machine Learning on Heterogeneous Systems. 2015. Available online: https://www.tensorflow.org (accessed on 28 May 2021).
  57. Chollet, F. Keras. 2015. Available online: https://keras.io (accessed on 28 May 2021).
  58. Kingma, D.P.; Ba, J. Adam: A Method for Stochastic Optimization. arXiv 2017, arXiv:1412.6980. Available online: https://arxiv.org/abs/1412.6980 (accessed on 28 May 2021).
  59. Srivastava, N.; Hinton, G.; Krizhevsky, A.; Sutskever, I.; Salakhutdinov, R. Dropout: A Simple Way to Prevent Neural Networks from Overfitting. J. Mach. Learn. Res. 2014, 15, 1929–1958. [Google Scholar]
  60. Hastie, T.; Tibshirani, R.; Friedman, J. The Elements of Statistical Learning: Prediction, Inference and Data Mining, 2nd ed.; Springer: New York, NY, USA, 2009. [Google Scholar]
  61. Okajima, T.; Sumitani, K.; Kawamoto, M.; Kobayashi, E. X-ray absorption spectroscopy using BL11 at SAGA-LS, and its applications for materials science. J. Phys. Conf. Ser. 2013, 430, 012088. [Google Scholar] [CrossRef] [Green Version]
  62. Rehr, J.J.; Albers, R.C. Theoretical approaches to X-ray absorption fine structure. Rev. Mod. Phys. 2000, 72, 621–654. [Google Scholar] [CrossRef]
Figure 1. Performance of the trained NN on the test dataset. On the top left, the MSE of the worst 200 cases are shown in descending order. Predicted RDF and BAD (insets) from selected configurations are shown in blue and the shadowed area represent standard deviation of the predictions. Orange lines are the same quantities calculated from the models, which are: (a) dia, (b) hcp, (c) fcc, (d) liquid and (e) bcc.
Figure 1. Performance of the trained NN on the test dataset. On the top left, the MSE of the worst 200 cases are shown in descending order. Predicted RDF and BAD (insets) from selected configurations are shown in blue and the shadowed area represent standard deviation of the predictions. Orange lines are the same quantities calculated from the models, which are: (a) dia, (b) hcp, (c) fcc, (d) liquid and (e) bcc.
Symmetry 13 01070 g001
Figure 2. Extracted k 2 -weighted EXAFS signal of Ni foil at room temperature.
Figure 2. Extracted k 2 -weighted EXAFS signal of Ni foil at room temperature.
Symmetry 13 01070 g002
Figure 3. RDF for Ni at 300 K obtained by MD simulation and by analysis of experimental data using NN. For comparison, peak position from X-ray diffraction [49] and RDF from RMC [42] are shown.
Figure 3. RDF for Ni at 300 K obtained by MD simulation and by analysis of experimental data using NN. For comparison, peak position from X-ray diffraction [49] and RDF from RMC [42] are shown.
Symmetry 13 01070 g003
Figure 4. Bond-angle distribution for Ni at 300K obtained by MD simulation and by analysis of experimental data using NN. Gray bars are fcc angles scaled to their coordination. Black points are the integral values of each peak for the NN prediction.
Figure 4. Bond-angle distribution for Ni at 300K obtained by MD simulation and by analysis of experimental data using NN. Gray bars are fcc angles scaled to their coordination. Black points are the integral values of each peak for the NN prediction.
Symmetry 13 01070 g004
Figure 5. RDF for undercooled and liquid Ni obtained by NN and compared with MD simulation and RMC analysis. On the right, the first peak of the RDF obtained by NN for the two temperatures are compared.
Figure 5. RDF for undercooled and liquid Ni obtained by NN and compared with MD simulation and RMC analysis. On the right, the first peak of the RDF obtained by NN for the two temperatures are compared.
Symmetry 13 01070 g005
Figure 6. Normalized BAD for undercooled and liquid Ni obtained by NN and compared with MD simulation and RMC analysis. On the right, the distributions from NN and RMC are compared.
Figure 6. Normalized BAD for undercooled and liquid Ni obtained by NN and compared with MD simulation and RMC analysis. On the right, the distributions from NN and RMC are compared.
Symmetry 13 01070 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Iesari, F.; Setoyama, H.; Okajima, T. Extracting Local Symmetry of Mono-Atomic Systems from Extended X-ray Absorption Fine Structure Using Deep Neural Networks. Symmetry 2021, 13, 1070. https://doi.org/10.3390/sym13061070

AMA Style

Iesari F, Setoyama H, Okajima T. Extracting Local Symmetry of Mono-Atomic Systems from Extended X-ray Absorption Fine Structure Using Deep Neural Networks. Symmetry. 2021; 13(6):1070. https://doi.org/10.3390/sym13061070

Chicago/Turabian Style

Iesari, Fabio, Hiroyuki Setoyama, and Toshihiro Okajima. 2021. "Extracting Local Symmetry of Mono-Atomic Systems from Extended X-ray Absorption Fine Structure Using Deep Neural Networks" Symmetry 13, no. 6: 1070. https://doi.org/10.3390/sym13061070

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop