Abstrct
Chemically disordered materials are widely utilized, yet establishing structure-property relationship remains challenging due to their vast configurational space. Identifying thermal accessible low energy configurations of these materials through standard ab initio calculations is computationally expensive for doping induced structure changes. In this work, we propose a straightforward algorithm to optimize random structures into ground state configurations by matching chemical subgraphs. This algorithm constructs harmonic potential with chemistry-driven parameterization, without relying on iterative training to accelerate the relaxation process. It can completely bypass the need for relaxation with ab initio calculations in rigid systems and reduce computational costs by 30% in flexible systems. Leveraging its exceptional structural relaxation capabilities, we have also developed a generalized workflow for screening low-energy structures in disordered materials, aimed at expediting the screening process and accelerating new material discovery.
Similar content being viewed by others
Introduction
In the realm of materials science, structural disorder, arising from impurities, defects1,2 or random atomic placements in high-entropy alloys3, enhances the diversity of material properties. While introducing structural disorder is a common strategy to modify material properties, it complicates the precise structure determination, both experimentally and theoretically4,5,6. Experimental techniques for probing disordered materials face limitations. X-ray diffraction (XRD) reveals long-range order but lacks local atomistic details7. Solid-state Nuclear Magnetic Resonance (NMR) resolves atomic positions and interactions but is limited to certain atoms with NMR activity8. Mössbauer spectroscopy (MES) offers detailed insights into atomic positions and coordination but is restricted to elements like Fe and Ni and suffers from temperature sensitivity9,10. These constraints hinder direct correlations between properties and atomic structures.
Theoretical approaches also face challenges in the discovery of disordered materials. First, chemical doping results in numerous nearly degenerate configurations existing in the thermodynamic limit. Identifying these low-energy configurations is essential for the high-throughput screening workflow. Second, chemical doping often causes structural changes that complicate property analysis and prediction due to reduced symmetry and more complex electronic structure calculations. Numerous computational methods have been developed for disordered materials. Early methods like the Virtual Crystal Approximation (VCA)11 and Coherent Potential Approximation (CPA)12 treated disordered structures as ideal crystals, overlooking doping induced lattice distortion and local atomic changes. The Special Quasirandom Structure (SQS) method13, introduced in 1990, improved accuracy by simulating short-range correlations but struggled with efficiency for complex systems. Structure enumeration programs, like Site Occupancy Disorder (SOD)14, enumlib15, Supercell16, and disorder17, generate all possible non-redundant structures but are computationally intensive with first-principles methods. The classic Cluster Expansion (CE) method is a well-established technique for efficiently calculating energies of various configurations in disordered materials18,19,20. More recently, the CE framework has been extended to incorporate force constants, enabling the prediction of thermodynamic properties, phonon density of states (DOS), and phase diagrams21,22,23. While, as highlighted by Nguyen et al.24, CE is effective for predicting ground-state energies when atomic relaxations are minimal, its predictive accuracy can be significantly diminished in systems exhibiting substantial atomic relaxations arising from chemical disorder. Ricardo et al.25 highlighted a strong correlation between electrostatic energy (E_elec) and ground state energy (E_opt), though E_elec is computation-free and widely used in high throughput screening, it may fail in non-ionic systems26,27,28. Fu et al.29 found that the single-point energy (E_sp) can reproduce E_opt results for systems with mild lattice relaxation. To develop an effective screening workflow, incorporating structural relaxation is essential. The main computational challenge is the costly structural relaxation of each initial structure in a large-scale dataset using density functional theory (DFT) calculations.
In recent years, many Machine Learning Potentials (MLPs)30,31,32,33 have been developed to predict the ground state structure from the initial configuration, known as the initial structure to relaxed structure (IS2RS) task34. These MLPs can significantly reduce computational costs by bypassing the need for DFT calculations to facilitate materials discovery. However, for disordered inorganic structures with varying compositions, MLPs face substantial challenges due to combinatorial explosion caused by doping. Ideally, they require separate training for each compositional space, which imposes prohibitively high data requirements. Active learning-based methods have been developed to alleviate this issue32,33,35,36,37, but the fundamental need to approximate the PES across diverse compositions remains a bottleneck. Furthermore, MLP-based relaxation is inherently indirect, as it approximates the PES before performing energy minimization. This process introduces error accumulation and potential deviations from true relaxed structures, particularly when generalizing to novel systems38. To address these limitations, recent efforts have explored end-to-end methods38,39,40,41,42. These approaches bypass MLP training entirely and directly predict relaxed structures from initial configurations, offering a promising solution for structural prediction in complex disordered systems. Yang et al.41,42 proposed DeepRelax, an iteration-free deep generative model using a periodicity-aware equivariant graph neural network (PaEGNN) with uncertainty quantification, enabling efficient and robust structural relaxation across diverse systems for scalable material discovery. Leveraging the power of graph convolutional neural networks (GCN), Kim and Zuo et al.39,40 applied the pix2pix domain translation model and BOWSR algorithm based on structure features to relax geometry without DFT. While achieving remarkable performance, they still require extensive DFT data for training and are typically material-specific43. Most notably, Yoon et al.38 used graph neural network (GNN) to develop DOGSS method, a machine-learned harmonic force field that approximates the ground state structure and properties of inorganic multicomponent surfaces. This approach eliminates the need for expensive electronic data such as energies, forces, and/or stresses, but it still requires improvement to address its dependence on extensive input structures.
To the best of our knowledge, the current mainstream end-to-end approaches are generally based on deep learning frameworks. While these methods are highly effective once established, their development is often requiring huge computing resources due to their data-hungry and black-box nature. Consequently, their applications in novel complex systems outside established databases like the Materials Project, is relatively rare for the out-of-distribution learning issue (i.e., a distribution different from the training distribution)44.
Inspired by the DOGSS, we propose a simple, chemistry-driven approach called the Structure Beautification Algorithm (SBA), which could predict ground state structure from initial configuration using surrogate harmonic potential directly constructed from a small dataset with chemistry-driven parameterization. As tested in rigid systems such as FeCo2SixAl1-x and ZnxCd1-xS, as well as the flexible iron carbide system, SBA has proven to be data-efficient, interpretable, and intuitive for approximating ground state structures. Building on these features, we establish an efficient high-throughput screening workflow for chemically disordered materials with extensive configurational space, accelerating material discovery.
Results and disscussion
Rigid structure systems
Heusler alloys are attractive for electromagnetic applications due to their unique properties, which are significantly influenced by atomic disorder45,46,47,48,49,50,51,52. Researches have concentrated on studying stable configurations and their properties in FeCo2SixAl1−x3,53,54,55,56,57,58,59. However, chemical disorder complicates synthesis and characterization. Similarly, in ZnxCd1-xS compounds, varying Zn content allows precise control over bandgap and optical properties, though it introduces computational challenges due to disordered structures60,61,62,63,64,65,66. Chemical disorder complicates experimental characterization. While theoretical studies focus on the most stable structures67, metastable ones with favorable properties may also exist68. Energy alone doesn’t determine the most relevant structure, statistical averaging with Boltzmann weights69, as shown in Supplementary Fig. S1, offers a more accurate analysis. A comprehensive understanding of disordered materials requires considering all low-energy configurations.
E_elec and E_sp are common descriptors for structure screening, and E_SBA_sp represents the single-point energy after SBA relaxation. We compare the ground state energy (E_opt) of each structure with E_elec and E_sp (Fig. 1). Pearson coefficients70 were computed based on Eq. (1) in the Supporting Information for E_elec/E_opt, E_sp/E_opt, and E_SBA_sp/E_opt. A coefficient of 1 indicates a perfect correlation, while a coefficient of -1 signifies a perfect anti-correlation. For FeCo2Si0.5Al0.5, the correlation coefficients are 82.19% for E_elec/E_opt and 90.37% for E_sp/E_opt, while E_SBA_sp/E_opt reaches 99.36% (Fig. 1a). In the Zn0.15Cd0.85S system, E_elec fails due to the ionic nature and E_sp shows a lower correlation of 54.19% with E_opt. In contrast, E_SBA_sp/E_opt is improved to 91.92% (Fig. 1b). Higher energy correlation of E_SBA_sp/E_opt enhances the search for more accurate low-energy structures, and improves the E_sp approach for structure screening.
a Comparative energy analysis of different methods in FeCo2Si0.5Al0.5 system. The left y-axis represents energy range of ground state energy (E_opt), single-point energy (E_sp) and the single-point energy after SBA relaxation (E_SBA_sp), while the right y-axis represents the electrostatic energy (E_elec). E_opt (gray circles), E_sp (red circles), E_SBA_sp (green circles) and E_elec (bule circles). b Comparison of energy profiles across various approaches for Zn0.15Cd0.85S. The left y-axis represents energy range of E_opt (gray circles) and E_SBA_sp (green circles), while the right y-axis represents the E_sp range (red circles).
In order to investigate the efficiency of the screening process, we also compare the cost-effectiveness of different methods. Here, we use Receiver Operating Characteristic Curve (ROC) to evaluate these methods in predicting energetically favorable configurations. ROC offers a comprehensive perspective on the performance of different classifier and exhibits robustness when changing the distribution of test samples data. To construct ROC, we define a true positive rate (TPR) as a stable structure predicted as stable, a false positive rate (FPR) as an unstable structure predicted as stable. We define 30 low-energy structures as target from E_opt rankings in the FeCo2Si0.5Al0.5 system and evaluate the performance of various methods (Fig. 2a). The ROC curve’s proximity to the top left corner indicates more effective classification of true low-energy structures. The area under the ROC curve (AUC) quantifies this, with values approaching 1 signifying superior performance. With an AUC of 0.99, single-point energy calculation after SBA (SBA+sp) is highly desirable to classify all stable structures for FeCo2Si0.5Al0.5 system. In contrast, single-point (sp) and electrostatic energy (elec) methods, with AUCs of 0.81 and 0.87 respectively, show less satisfactory performance. Moreover, the computational cost analysis (Fig. 2b) supports this, showing that the SBA+sp method dramatically cut total costs to 31.94 and waste to 0.94 h, compared to 80 and 50 h for elec and sp, respectively. This underscores its superior performance in reducing computational overhead. Trace back to the root cause, we find that SBA significantly reduces the per atom force from 0.64 eV/Å to 0.25 eV/Å, yielding more accurate geometries from the initial configurations. This improvement leads to higher accuracy and efficiency.
a Receiver operating characteristic curve for low energy configurations prediction. It displays the ability of electrostatic energy (elec), single-point energy (sp), and SBA calculations to classify stable structures when applied to initial structures. The x-axis is the fraction of unstable structures classified as stable. The y-axis is the fraction of stable structures classified as stable. The area under curve (AUC) is also shown for all methods. b Sampling cost for screening low-energy structures in different methods. Total-cost (gray-blue bar) is the overall cost to screen target low-energy structures, while Waste-cost (gray-orange bar) represents the extra computational expense incurred due to inaccuracies in different methods.
Benefiting from the exceptional structure prediction ability of the SBA, we establish a generalized workflow for screening low-energy structures in disordered materials (Fig. 3), different systems undergo distinct processes. For rigid systems, we initially apply SBA to relax initial structures, followed by a straightforward sp calculation and bypassing time-cost DFT procedure, allowing for the efficient identification of low-energy structures. Since SBA advances the value of sp calculation and this workflow may have the potential to replace high-throughput screening in such systems.
To validate the rationality of relaxed structure by SBA, we compare the electronic properties of FeCo2Si0.5Al0.5 for initial, ground state, and SBA-relaxed configurations. Figure 4a shows all three structures exhibit a consistent spin-down band gap of 0.70 eV near the Fermi level, indicating half-metallic behavior. Pearson coefficients of 94.8% and 99.9% for initial and SBA-relaxed structures compared to the ground state demonstrate accuracy structure prediction ability of SBA. For 10 randomly selected configurations, SBA relaxation reduces discrepancies in the averaged band gap and magnetic moment, aligning properties closer to the ground state (Fig. 4b). Likewise, for Zn0.15Cd0.85S, the average band gap deviation is just 1.5% (Fig. 4c), and the DOS similarity between SBA-relaxed and ground state structures is 96.54% (Fig. 4d). This remarkable accuracy allows SBA to not only replace high-throughput calculations for screening low-energy structures but also to predict electronic properties directly, achieving over 95% savings in property evaluation costs. CE methods predict thermodynamic properties using a limited set of first-principles calculations but fail to explicitly account for lattice relaxation effects21,22,23. In contrast, SBA incorporates relaxation into harmonic potential construction, generating optimized structural analogs to mitigate lattice relaxation impact. SBA is particularly suited for systems requiring explicit relaxation treatment, while CE remains advantageous for large-scale configurational sampling due to its lower computational overhead. To more efficiently and comprehensively demonstrate the robustness and universality of the SBA, we use the Spanish Initiative for Electronic Simulations with Thousands of Atoms code (SIESTA) program71 which offers lower computational cost, to perform calculations on the Ni1-xMox (x = 0.4) alloy system, supplementary Fig. S2–S3 further highlight the excellent performance of SBA.
a Comparison of the DOS for the initial structure (light-gray shading), SBA-relaxed structure (light-bule shading) and ground state structure (light-purple shading) for FeCo2Si0.5Al0.5 system. b The band gap and magnetic moment difference for the initial structures and SBA-relaxed structures for FeCo2Si0.5Al0.5 system. The left y-axis stands band gap (bule bar) and right y-axis is magnetic moment difference (orange bar). c Comparison of the DOS for the initial structure (light-gray shading), SBA-relaxed structure (light bule shading) and ground state structure (light- purple shading) in the Zn0.15Cd0.85S system. d Comparison of the average band gap for the initial structure (light-purple bar), SBA-relaxed structure (gray-blue bar), and ground state structure (gray-orange bar) in the Zn0.15Cd0.85S system.
Flexible structure with floppy potential energy surface
Iron carbide intermetallic compounds play a crucial role in catalysis for the carbon nanotubes (CNTs)72, pollutant removal73, electro-catalysis74, electronic water splitting and Fischer-Tropsch synthesis. It is well known that the catalytic activities are mainly determined by their structures75,76, which strongly affects the number of active sites and the electronic structure of heterogeneous catalysts. Discerning the active structure is significant to establish the intrinsic structure-activity relationship, and it is the basis for high-performance catalysts design. However, iron and carbon could form variable disordered structures during the reaction conditions, which brings a challenge for experimental characterization. Therefore, numerous studies have been carried out to improve the catalytic performance by designing well-defined structures. In theoretical studies, Gao et al.77. used molecular dynamics to analyze carburization in Body-Centered Cubic (BCC) iron and found lattice changes to Face-Centered Cubic (FCC) and Hexagonal Close-Packed (HCP) in late stage. Yuan et al.78 systematically explored iron carbide phases through the global structure search approach coupled with DFT methods and enriched the understanding of local structure and properties of iron carbides.
In general, the iron carbide catalysts are usually synthesized through iron carburization by carbon-source feedstock. Accordingly, in this work, the FCC iron lattice serves as the host, and the effect of the tiny content-permeated carbon atoms located at the octahedral site79 on the iron structure is investigated. Enumeration and Boltzmann distribution weight (Fig. 5a) are employed to derive reasonable low-energy structures, with Pearson coefficient evaluating the correlation between E_elec/E_opt, E_sp/E_opt. Correlations of 46.19% and 50.3% (Fig. 5b) indicate that both E_sp and E_elec are inadequate for reliably screening stable candidates in this flexible system. Although the SBA method is applied to relax the initial structures, the correlation of E_SBA_sp/E_opt only improves slightly to 54.2%. This poor correlation is attributed to the mismatch in Fe/C atomic sizes in the C-doped system. However, the averaged per atom force within the structure decreases from 2.77 eV/Å to 1.34 eV/Å, which indicates that the structures become more reasonable in geometry after SBA relaxation. Further speaking, when performing full relaxation based on SBA-relaxed geometries, different initial structures exhibit varying degrees of computational resource savings, with maximum savings reaching up to 60% (Fig. 6a). The ionic steps and time required to reach a convergent state are both reduced by 33% (Fig. 6b). These results demonstrate the impact of SBA in accelerating calculations for flexible systems. Supplementary Figs. S4–S5 and Table S1, derived from SIESTA calculations for the Al2.5CoNi4.5 system, provide additional evidence supporting the performance of SBA.
a Boltzmann proportion of low-energy configurations under specific temperature condition. b Energy landscape of different computational approaches in the Fe32C4 System. The energy landscape for various approaches, including ground state energy (E_opt), single-point energy (E_sp), single-point energy after SBA relaxation (E_SBA_sp) and electrostatic energy (E_elec). The left y-axis represents the energy range of E_opt, E_sp, and E_SBA_sp, while the right y-axis shows the range of E_elec. E_opt (gray circles), E_sp (red circles), E_elec (blue circles), and E_SBA_sp (green circles).
a Time cost improvement of each configuration after SBA relaxation. Time cost improvement refers to the savings in computational resources compared to DFT structure optimization. b The comparison of structural optimization convergence rates. The left y-axis stands for ionic steps and right y-axis is time cost of DFT relaxation (gray-purple bar) and SBA-relaxed relaxation (gray-orange bar).
A cost-effective method, i.e., E_elec approach, is valuable to reduce the structural space when dealing with flexible system with vast configuration space. Although the accuracy of E_elec method is limited, it is highly effective for rapid structure screening and has received positive feedback in the literatures26,27,28. Alternatively, there is an advanced sampling space reduction algorithms to speed up the structural screening process, i.e., LAsou technique, developed by our group. It is based on machine learning potentials but proved its efficiency in many material fields36,80. By combining LAsou with SBA, the process of structural screening would become more simplified and manageable in the future.
In summary, we introduce an algorithm that leverages the locality of chemical environments and develop a parameters-free harmonic potential for refining the unreasonable initial structures. This algorithm represents a significant advancement, moving beyond the prevailing reliance on data-driven MLPs for predicting stable structures. For rigid systems, SBA attains over 90% agreement in energy between SBA-relaxed and ground state structures, effectively identifying low-energy configurations and reducing property calculation costs by more than 95%. Even for the challenging flexible systems, it achieves a 30% reduction in computational cost, comparable performance to the DOGSS model. Since the SBA method does not depend on large training dataset, it exhibits notable transferability and practicality. By employing its advanced relaxation capability, we establish a comprehensive workflow for the systematic identification of low-energy structures in disordered materials. This methodology aims to improve the efficiency of the screening process and accelerate the discovery of novel materials.
Methods
Structure beautification algorithm
As illustrated in Fig. 7, algorithms for predicting ground state structures from input initial structures can be roughly classified into three categories. MLPs are typically constructed using computationally expensive DFT data to fit the global potential energy surface (PES), enabling predictions of ground-state configurations from structural inputs (Fig. 7a). While active learning strategies33,34,36,37,80,81,82 improve data efficiency, unresolved challenges remain. In complex systems like disordered doping, the vast configuration space severely limits active learning efficacy for limited transferability across different compositional spaces. To address this practical issue, Yoon et al.38 introduced an innovative approach to accelerate structure relaxation of inorganic multicomponent surface. This approach avoids the need for expensive datasets containing energies and force, only relying on initial and relaxed structure pairs. Rather than directly constructing a global PES, DOGSS first predicts the parameters for a local PES of a given input structure using a GNN model, represented by classic model potential, such as harmonic or L-J potential. The predicted local PES is then minimized to yield relaxed structure (Fig. 7b). This model is not evaluated on how well it fits energies and forces, but on how well the predicted structure matches the ground state structure. It greatly simplifies data preparation, achieving better transferability, wider applicability and practical utility38. Nevertheless, DOGSS is constructed on a deep learning framework within GNN, it necessitates a substantially large training dataset (GASPy) for initialization.
a The overall workflow to utilize MLPs for ground structure. b Workflow of the DOGSS Algorithm38. c Proposed Workflow in this work.
Follow the direction innovated by Yoon38, we are exploring the possibility of simplifying dataset preparation by demystifying the black box of machine learning part with chemical insights. As illustrated in Fig. 7c, we propose a straightforward chemistry-driven model to predict the ground state of input structure. This approach circumvents the traditional necessity for extensive, high-dimensional training datasets, offering a more efficient pathway for high-throughput screening of novel complex disordered structures. The fundamental chemical insight behind SBA is that the complex configurational space can be significantly simplified at the coarse-grained local chemical environment level using graph isomorphism. Graph theory is widely applied in areas such as data science and chemical engineering83,84. Greeley et al.85 applied a graph theory-based approach to simplify surface adsorption complexities, facilitating the identification of unique configurations and the systematic estimation of high coverage models on low-symmetry catalytic surfaces.
In our method, each atom within a disordered configuration is characterized by a subgraph representing its local chemical environment, comprising the central atom and its nearest neighbors within a defined cutoff, as illustrated in Table 1. These subgraphs, represented as undirected graphs (atoms as nodes, bonds as edges), are then subjected to redundancy filtering to identify unique topological motifs for reference. For simplicity, duplicate motifs are randomly represented by one subgraph. Similar to DOGSS, SBA constructs a local harmonic potential for a given structure and then minimized it to approximate the ground state structure. In contrast to traditional machine learning potentials requiring extensive iterative training, our harmonic potential parameters are directly derived from a very small pre-computed dataset of relaxed structures without iterative fitting. The necessary subgraphs for assembling the topological structure are identified within the reference subgraphs, enabling the direct retrieval of the pairwise parameters for optimal spring distances lij in harmonic potential. For each pair of atom i and atom j, the required spring constant parameter is set to k = 1/lij2, following the force-directed graph drawing algorithm86. Hence, the local harmonic potential is constructed to approximate the ground state spatial structure without iterative fitting.
Besides energies and atomic forces of harmonic potential, the viral stress inside a periodic cell is derived as formulated by Thompson et al.87 to deal with variable lattice issues. The energy minimization was carried out using the FIRE optimizer88 from the Atomic Simulation Environment (ASE) package89. A stringent force convergence criterion of 0.001 eV/Å. During this process, atomic positions were relaxed until the maximum force on any atom decreases below the convergence criterion.
Computational details
To evaluate the feasibility of SBA as illustrated in Fig. 7c, we demonstrated the performance of our model on both rigid and flexible systems. In the context of chemical doping, rigid systems exhibit minor deviation from initial structures, whereas flexible systems can result in significant lattice distortions. Based on these system characteristics, different strategies can be developed for high-throughput screening. Specially, we used FeCo2SixAl1-x (x = 0.5) and ZnxCd1-xS (x = 0.15) systems as cases for rigid systems, and iron-carbide Fe32Cx (x = 4) for flexible systems. All initial structures were generated using the open-source Supercell program16, allowing for the customization of supercell sizes and lattice site occupancy. 153, 325, and 71 structures were generated for FeCo2SixAl1-x (x = 0.5), ZnxCd1-xS (x = 0.15), and Fe32Cx (x = 4) systems, respectively.
We performed first-principles calculations using the plane wave code Vienna ab initio simulation package (VASP)90,91, the generalized gradient approximation of the Perdew-Burke-Ernzerhof parameterization (GGA-PBE)92 was adopted for the exchange and correlation functions. In addition, the PBE + U method93 (Ueff = 2 eV for Co and Fe)56 was employed. All atoms were allowed to relax to their equilibrium states, with an energy cutoff set of 500 eV and convergence criteria of 1 × 10–4 eV for energy and 0.02 eV/Å for forces.
Data availability
All data needed to evaluate the conclusions in the paper are present in the paper and/or the Supplementary Information, and available from the corresponding authors upon reasonable request.
Code availability
The code Structure Beautification Algorithm in this paper is publicly available at https://github.com/spdkit/SI_SBA.
References
Kohiki, S., Nishitani, M., Wada, T. & Hirao, T. Enhanced conductivity of zinc oxide thin films by ion implantation of hydrogen atoms. Appl. Phys. Lett. 64, 2876–2878 (1994).
Van de Walle, C. G. Hydrogen as a cause of doping in zinc oxide. Phys. Rev. Lett. 85, 1012–1015 (2000).
Inomata, K. et al. Site disorder in Co2Fe(Al,Si) Heusler alloys and its influence on junction tunnel magnetoresistance. Phys. Rev. B. 77, 214425 (2008).
Grau-Crespo, R., de Leeuw, N. H. & Catlow, C. R. A. Cation distribution and magnetic ordering in FeSbO4. J. Mater. Chem. 13, 2848–2850 (2003).
Soven, P. Contribution to the theory of disordered alloys. Phys. Rev. 178, 1136–1144 (1969).
Yang, Y. et al. An unusual strong visible‐light absorption band in red anatase TiO2 photocatalyst induced by atomic hydrogen‐occupied oxygen vacancies. Adv. Mater. 30, 1704479 (2018).
Li, J. & Sun, J. Application of X-ray diffraction and electron crystallography for solving complex structure problems. Acc. Chem. Res. 50, 2737–2745 (2017).
Bryce, D. L. NMR crystallography: structure and properties of materials from solid-state nuclear magnetic resonance observables. IUCrJ 4.4, 350–359 (2017).
Nakamura, S. et al. Crystal-site-selective spectrum of Fe3O4 obtained by mössbauer diffraction. J. Phys. Soc. Jpn. 86, 023706 (2017).
Ounacer, M. et al. Structural, magnetic, and Mössbauer studies of magnetite and nickel-copper and nickel-zinc ferrites. Inorg. Chem. Commun. 158, 111650 (2023).
Soderlind, P. et al. Spin and orbital magnetism in Fe-Co and Co-Ni alloys. Phys. Rev. B 45, 12911–12916 (1992).
Soven, P. Coherent-potential model of substitutional disordered alloys. Phys. Rev. 156, 809–813 (1967).
Zunger, A., Wei, S., Ferreira, L. G. & Bernard, J. E. Special quasirandom structures. Phys. Rev. Lett. 65, 353–356 (1990).
Grau-Crespo, R., Hamad, S., Catlow, C. R. A. & Leeuw, N. H. d. Symmetry-adapted configurational modelling of fractional site occupancy in solids. J. Phys. Condens. Matter 19, 256201 (2007).
Hart, G. L. W. & Forcade, R. W. Algorithm for generating derivative structures. Phys. Rev. B. 77, 22415 (2008).
Okhotnikov, K., Charpentier, T. & Cadars, S. Supercell program: a combinatorial structure-generation approach for the local-level modeling of atomic substitutions and partial occupancies in crystals. J. Cheminf. 8, 1–15 (2016).
Lian, J.-C. et al. Algorithm for generating irreducible site-occupancy configurations. Phys. Rev. B. 102, 134209 (2020).
Wu, Q. et al. Cluster expansion method and its application in computational materials science. Comput. Mater. Sci. 125, 243–254 (2016).
Schulthess, T., Monnier, R. & Crampin, S. Effective cluster interactions at alloy surfaces and charge self-consistency: Surface segregation in Ni-10 at. % Al and Cu-Ni. Phys. Rev. B 50, 18564–18571 (1994).
Hart, G. L., Blum, V., Walorski, M. J. & Zunger, A. Evolutionary approach for determining first-principles hamiltonians. Nat. Mater. 4, 391–394 (2005).
Van De Walle, A. & Ceder, G. J. Ro. M. P. The effect of lattice vibrations on substitutional alloy thermodynamics. Rev. Mod. Phys. 74, 11 (2002).
Zarkevich, N. A. & Johnson, D. D. J. P. r. l. Reliable first-principles alloy thermodynamics via truncated cluster expansions. Phys. Rev. Lett. 92, 255702 (2004).
Berthier, F., Lullien, Q. & Legrand, B. J. P. R. B. Effective site energy and cluster expansion approaches for the study of phase diagrams. Phys. Rev. B 104, 014111 (2021).
Nguyen, A. H., Rosenbrock, C. W., Reese, C. S. & Hart, G. L. W. Robustness of the cluster expansion: assessing the roles of relaxation and numerical error. Phys. Rev. B 96, 014107 (2017).
Grau-Crespo, R., de Leeuw, N. H. & Catlow, C. R. A. Distribution of cations in FeSbO4: a computer modeling study. Chem. Mater. 16, 1954–1960 (2004).
Mo, Y., Ong, S. P. & Ceder, G. First principles study of the Li10GeP2S12 lithium super ionic conductor material. Chem. Mater. 24, 15–17 (2011).
Qi, C. et al. Machine learning-accelerated first-principles study of atomic configuration and ionic diffusion in Li10GeP2S12 solid electrolyte. Materials 17, 1810 (2024).
Oh, K. et al. Native defects in Li10GeP2S12 and their effect on lithium diffusion. Chem. Mater. 30, 4995–5004 (2018).
Fu, G., Xu, X. & Sautet, P. Vanadium distribution in four-component Mo-V-Te-Nb mixed-oxide catalysts from first principles: how to explore the numerous configurations? Angew. Chem. Int. Ed. 51, 12854–12858 (2012).
Xie, S. R., Rupp, M. & Hennig, R. G. Ultra-fast interpretable machine-learning potentials. npj Comput. Mater. 9, 162 (2023).
Lemm, D., von Rudorff, G. F. & von Lilienfeld, O. A. Machine learning based energy-free structure predictions of molecules, transition states, and solids. Nat. Commun. 12, 4468 (2021).
Garijo del Río, E., Mortensen, J. J. & Jacobsen, K. W. J. P. R. B. Local Bayesian optimizer for atomic structures. Phys. Rev. B. 100, 104103 (2019).
Bisbo, M. K. & Hammer, B. J. P. R. B. Global optimization of atomic structure enhanced by machine learning. Phys. Rev. B. 105, 245404 (2022).
Chanussot, L. et al. Open Catalyst 2020 (OC20) dataset and community challenges. ACS Catal. 11, 6059–6072 (2021).
Vandermause, J. et al. Active learning of reactive Bayesian force fields applied to heterogeneous catalysis dynamics of H/Pt. Nat. Commun. 13, 5183 (2022).
Yuan, X. et al. Active learning to overcome exponential-wall problem for effective structure prediction of chemical-disordered materials. npj Comput. Mater. 9, 12 (2023).
Ma, H. et al. Machine learning predicts atomistic structures of multielement solid surfaces for heterogeneous catalysts in variable environments. Innovation 5, 100571 (2024).
Yoon, J. & Ulissi, Z. W. Differentiable Optimization for the Prediction of Ground State Structures (DOGSS). Phys. Rev. Lett. 125, 173001 (2020).
Kim, S. et al. A structure translation model for crystal compounds. npj Comput. Mater. 9, 142 (2023).
Zuo, Y. et al. Accelerating materials discovery with Bayesian optimization and graph deep learning. Mater. Today 51, 126–135 (2021).
Yang, Z. et al. Scalable crystal structure relaxation using an iteration-free deep generative model with uncertainty quantification. Nat. Commun. 15, 8148 (2024).
Yang, Z. et al. Scaling crystal structure relaxation with a universal trustworthy deep generative model. Condens. Matter 2404, 00865 (2024).
Kocer, E., Ko, T. W. & Behler, J. Neural network potentials: a concise overview of methods. Annu. Rev. Phys. Chem. 73, 163–186 (2022).
Hu, J., Liu, D., Fu, N. & Dong, R. Realistic material property prediction using domain adaptation based machine learning. Digit Discov. 3, 300–312 (2024).
Inomata, K. et al. Highly spin-polarized materials and devices for spintronics. Sci. Technol. Adv. Mater. 9, 014101 (2008).
Kc, S. et al. Co2Fe1.25Ge0.75: a single-phase full Heusler alloy with highest magnetic moment and Curie temperature. Acta Mater. 236, 118112 (2022).
Basit, L. et al. Quaternary Heusler compounds without inversion symmetry: CoFe1+xTi1–xAl and CoMn1+xV1–xAl. Eur. J. Inorg. Chem. 2011, 3950–3954 (2011).
Gao, Q., Opahle, I. & Zhang, H. High-throughput screening for spin-gapless semiconductors in quaternary Heusler compounds. Phys. Rev. Mater. 3, 024410 (2019).
Trudel, S., Gaier, O., Hamrle, J. & Hillebrands, B. Magnetic anisotropy, exchange and damping in cobalt-based full-Heusler compounds: an experimental review. J. Phys. D Appl. Phys. 43, 193001 (2010).
Ksenofontov, V. et al. Hyperfine magnetic field on iron atoms and Co–Fe disordering in Co2FeSi. J. Appl. Phys. 107, 668 (2010).
Tezuka, N. et al. Tunnel magnetoresistance for junctions with epitaxial full-Heusler Co2FeAl0.5Si0.5 electrodes with B2 and L21 structures. Appl. Phys. Lett. 89, 112514 (2006).
Kc, S. et al. Tunable properties and potential half-metallicity in (Co2-xTix)FeGe Heusler alloys: An experimental and theoretical investigation. Phys. Rev. Mater. 3, 114406 (2019).
Fecher, G. H. & Felser, C. Substituting the main group element in cobalt–iron based Heusler alloys: Co2FeAl1-xSix. J. Phys. D: Appl. Phys. 40, 1582–1586 (2007).
Balke, B., Fecher, G. H. & Felser, C. Structural and magnetic properties of Co2FeAl1-xSix. Appl. Phys. Lett. 90, 242503 (2007).
Jorge, E. A. et al. Magnetic and structural properties of Co2FeAl1-xSix thin films. J. Phys. Conf. Ser. 200, 072006 (2010).
Yang, Y. et al. First-principles study of structure, electronic structure and thermoelectric properties for Co-based Heusler alloys Co2FeAl1-xSix (x = 0.25, x = 0.5, x = 0.75). Acta Phys. Sin. 68, 046101 (2019).
Nishihara, H. et al. Ab initio study of 59Co NMR spectra in Co2FeAl1-xSix Heusler alloys. Phys. B 465, 66–70 (2015).
Wurmehl, S., Kohlhepp, J. T., Swagten, H. J. M. & Koopmans, B. 59Co nuclear magnetic resonance study of the local distribution of atoms in the Heusler compound Co2FeAl0.5Si0.5. J. Appl. Phys. 111, 043903 (2012).
Wójcik, M. et al. 59Co NMR experiment as a probe of electron doping in Co2FeAl1−xSix Heusler alloys. Phys. Rev. B 85, 100401 (2012).
Bziz, I., Atmani, E. H., Fazouan, N. & Aazi, M. First-principles calculations of structural, electronic and optical properties of CdTexS1-x and Cd1-xZnxS ternary alloys. Surf. Interfaces 24, 101126 (2021).
Lilhare, D., Sinha, T. & Khare, A. Influence of Cu doping on optical properties of (Cd–Zn)S nanocrystalline thin films: a review. J. Mater. Sci. Mater. Electron. 29, 688–713 (2017).
Gaewdang, N. & Gaewdang, T. Investigations on chemically deposited Cd1−xZnxS thin films with low Zn content. Mater. Lett. 59, 3577–3584 (2005).
Azizi, S., Rezagholipour Dizaji, H. & Ehsani, M. H. Structural and optical properties of Cd1-xZnxS (x = 0, 0.4, 0.8 and 1) thin films prepared using the precursor obtained from microwave irradiation processes. Optik 127, 7104–7114 (2016).
Mochahari, P. K. & Sarma, K. C. Study of structural and optical properties of chemically synthesized nanostructured cadmium zinc sulphide films for band gap tunability. Indian J. Phys. 90, 21–27 (2015).
Nishidate, K. et al. Density-functional electronic structure calculations for native defects and Cu impurities in CdS. Phys. Rev. B 74, 035210 (2006).
Liu, X.-D. & Xing, T. First-principles calculations of electronic structures and optical properties of group-IIIA elements doped wurtzite CdS. Solid State Commun. 187, 72–76 (2014).
Wang, Y., Lv, J., Gao, P. & Ma, Y. Crystal structure prediction via efficient sampling of the potential energy surface. Acc. Chem. Res. 55, 2068–2076 (2022).
Aitchison, C. M. et al. Photocatalytic proton reduction by a computationally identified, molecular hydrogen-bonded framework. J. Mater. Chem. A 8, 7158–7170 (2020).
Zhou, Z., Shi, J., Wu, P. & Guo, L. Configuration dependence of the properties of Cd1–xZnxS solid solutions by first‐principles calculations. Phys. Status Solidi B 251, 655–660 (2014).
Zhang, W. Y., Wei, Z. W., Wang, B. H. & Han, X. P. Measuring mixing patterns in complex networks by Spearman rank correlation coefficient. Phys. A 451, 440–450 (2016).
Hirshman, S. P., Sanchez, R. & Cook, C. R. SIESTA: a scalable iterative equilibrium solver for toroidal applications. Phys. Plasma 18, 062504 (2011).
He, Z. et al. Iron Catalysts for the Growth of Carbon Nanofibers: Fe, Fe3C or Both? Chem. Mater. 23, 5379–5387 (2011).
Wang, S. et al. Biochar-supported nZVI (nZVI/BC) for contaminant removal from soil and water: a critical review. J. Hazard. Mater. 373, 820–834 (2019).
Song, A. et al. Uniform multilayer graphene-coated iron and iron-carbide as oxygen reduction catalyst. ACS Sustain. Chem. Eng. 6, 4890–4898 (2018).
Imaoka, T. et al. Platinum clusters with precise numbers of atoms for preparative-scale catalysis. Nat. Commun. 8, 688 (2017).
Cao, S. et al. Size-and shape-dependent catalytic performances of oxidation and reduction reactions on nanocatalysts. Chem. Soc. Rev. 45, 4747–4765 (2016).
Gao, R. et al. Carbon Permeation: the Prerequisite elementary step in iron-catalyzed Fischer–Tropsch Synthesis. Catal. Lett. 149, 645–664 (2019).
Yuan, X. et al. Crystal structure prediction approach to explore the iron carbide phases: novel crystal structures and unexpected magnetic properties. J. Phys. Chem. C. 124, 17244–17254 (2020).
Domain, C., Becquart, C. S. & Foct, J. Ab initiostudy of foreign interstitial atom (C, N) interactions with intrinsic point defects in α-Fe. Phys. Rev. B 69, 144112 (2004).
Peng, Q. et al. Active-learning search for unitcell structures: a case study on Mg3Bi2-xSbx. Comput. Mater. Sci. 226, 112260 (2023).
Perego, S. & Bonati, L. Data efficient machine learning potentials for modeling catalytic reactivity via active learning and enhanced sampling. npj Comput. Mater. 10, 291 (2024).
Chen, B. W., Zhang, X. & Zhang, J. J. C. S. Accelerating explicit solvent models of heterogeneous catalysts with machine learning interatomic potentials. Chem. Sci. 14, 8338–8354 (2023).
Kulkarni, J. & Humanities, S. Graph theory: applications to chemical engineering and chemistry. Galore Int. J. Appl. Sci. 1, 2017 (2017).
Prathik, A., Uma, K. & Anuradha, J. An overview of application of graph theory. Int. J. ChemTech Res. 9, 242–248 (2016).
Deshpande, S., Maxson, T. & Greeley, J. Graph theory approach to determine configurations of multidentate and high coverage adsorbates for heterogeneous catalysis. npj Comput. Mater. 6, 79 (2020).
Kamada, T. S. K. An algorithm for drawing general undirected graphs. Inf. Process. Lett. 31, 7–15 (1989).
Thompson, A. P., Plimpton, S. J. & Mattson, W. General formulation of pressure and stress tensor for arbitrary many-body interaction potentials under periodic boundary conditions. J. Chem. Phys. 131, 154107 (2009).
Bitzek, E. et al. Structural relaxation made simple. Phys. Rev. Lett. 97, 170201 (2006).
Larsen, A. H. et al. The atomic simulation environment—a Python library for working with atoms. J. Phys. Condens. Matter 29, 273002 (2017).
Kresse, G. & Furthmüller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 6, 15–50 (1996).
Kresse, G. & Hafner, J. Ab initio molecular dynamics for liquid metals. Phys. Rev. B 47, 558–561 (1993).
Perdew, J. P., Burke, K. & Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 77, 3865–3868 (1996).
Noorafshan, M. & Nourbakhsh, Z. The effect of Ce dilution on the ferromagnetic ordering and Kondo behavior of CeRuPO. J. Magn. Magn. Mater. 426, 287–293 (2017).
Acknowledgements
This work was financially supported by the National Science Fund for Distinguished Young Scholars of China (Grant No. 22225206), the National Key R&D Program of China (No. 2022YFA1604103), the National Natural Science Foundation of China (Nos. 22202224, 21972157), the CAS Project for Young Scientists in Basic Research (YSBR-005), the Key Research Program of Frontier Sciences CAS (ZDBS-LY-7007), the Major Research Plan of the National Natural Science Foundation of China (92045303), the Informatization Plan of the Chinese Academy of Sciences (Grant No. CAS-WX2021SF0110). Funding support was also received from the Beijing Advanced Innovation Center for Materials Genome Engineering, Synfuels China Co., Ltd., and the Institute of Coal Chemistry, Chinese Academy of Sciences.
Author information
Authors and Affiliations
Contributions
X.W., W.G. and J.L. designed research; L.A. and W.G performed research and developed the SBA Method; L.A., J.L., H.M. and W.G. analyzed data; L.A. and W.G. wrote the paper; J.L., W.G. and H.M. polished the language and revised the paper. All the authors contributed to the idea and reviewed the manuscript.
Corresponding authors
Ethics declarations
Competing interests
The authors declare no competing interests.
Additional information
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Supplementary information
Rights and permissions
Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
About this article
Cite this article
An, L., Ma, H., Liu, J. et al. Accelerating structure relaxation in chemically disordered materials with a chemistry-driven model. npj Comput Mater 11, 226 (2025). https://doi.org/10.1038/s41524-025-01694-3
Received:
Accepted:
Published:
DOI: https://doi.org/10.1038/s41524-025-01694-3