FT2DP: large atomic model fine-tuned machine learning potential for accelerating atomistic simulation of iron-based Fischer-Tropsch synthesis
Abstract
Density-functional theory (DFT)-based atomistic simulation methods have been essential in studying the structure-property relationships in heterogeneous catalysis. However, for complex catalytic processes, such as iron-based Fischer-Tropsch synthesis (FTS), the temporal or spatial scales involved are generally too large to perform DFT calculations. Recently, the development of machine learning potentials (MLPs) has demonstrated the capability for atomistic simulation on a large scale and long duration, and the rise of large atomic models (LAMs) is gaining much attention with unified descriptors incorporating a wide range of chemical knowledge and fine-tuning methodology for efficiently deploying the model to downstream tasks. In this work, we construct a MLP named fine-tuned Fischer-Tropsch deep potential (FT
Keywords
INTRODUCTION
As a key aspect of the chemical industry, heterogeneous catalysis has played a crucial role in the large-scale production of commodity chemicals such as ammonia, alcohol, and synthetic fuels. Nowadays, computational simulations based on ab initio density-functional theory (DFT) calculations have offered unprecedented opportunities for the rational design of novel solid catalysts by providing a deep atomistic analysis of the structures and reaction properties combined with theories in heterogeneous catalysis [1]. However, accurate and efficient computational simulations of complex heterogeneous catalytic processes remain highly challenging because of the demanding computational cost. One of the typical examples of complex heterogeneous catalytic systems is the Fischer-Tropsch synthesis (FTS), which converts syngas (a mixture of CO and H2) into fuels and other valuable chemicals, holding a special position in the energy industry [2, 3]. Among many possible candidates, the iron-based catalyst has gained much attention due to its low cost, high sulfur tolerance, and low methane selectivity [4, 5]. The active phases of iron-based FTS are believed to be in situ formed iron carbides such as
To achieve the goal of rational catalyst design, extensive research has been conducted to elucidate the relationship between the surface structures of iron carbides and their reactivities in the FTS process [14-18]. For instance, Chen et al. found that the CO dissociation barrier on different
The past several years have witnessed enormous advances in artificial intelligence (AI) methods, fueling a new paradigm shift of discoveries in natural sciences and giving rise to a new area of research, known as AI for science [19]. Especially, with the rapid development of machine learning potentials (MLPs) with both the accuracy of DFT and the efficiency of classical force fields, the detailed atomistic simulation for complex heterogeneous catalysis systems on a large temporal and spatial scale has become feasible[20], so that there emerge a large number of so-called AI-driven atomistic simulation platforms incorporating MLPs sharing the same basic architecture that using structural descriptors to represent the atomic structures and a fitting model (usually an artificial neural network) to decode the representation from descriptor to the potential energy. One notable example is the LASP software [21] with its neural network potential constructed by the power-type structural descriptors represented by atom-centered symmetry functions [22] developed by Huang et al., which has been widely utilized in computational simulation for various heterogeneous catalysis topics. For instance, in the ethene epoxidation reaction on silver[23], they identified the O
Despite numerous successes of MLPs in atomistic simulations, their applications often face economic and scalability limitations. The most obvious one is that all of the data used for training and validating MLPs are generated from scratch, namely, by ab initio molecular dynamic (AIMD) simulation or other simulation methods with ab initio calculation (typically DFT). Efficient data generation platforms through an active learning procedure such as DP GENerator (DP-GEN)[37] or Generating DP with Python (GDPy)[38] can significantly facilitate this process. However, a substantial amount of effort is still needed to construct DFT-labeled datasets, especially for heterogeneous catalysis domain due to its complexity with characteristics of bulks, surfaces, and adsorbates altogether. The publicly available large datasets, such as OC20[39], OC22[40], and MPtrj[41], have covered extensive physical and chemical knowledge, becoming very useful for overcoming the challenges of data generation. However, traditional MLPs usually struggle to generalize to applications not covered by the training data, especially when additional elements and structural configuration are included in the simulation tasks, making them poor at combining and utilizing the knowledge from these multiple and large datasets together due to the varying ab initio calculation methods employed in different datasets and the vast compositional space contained across all these datasets.
Recently, the rise of "universal" or "fundamental" MLPs offers opportunities for addressing the issues above and greatly extending the application scope of MLPs, often referred to as large atomic models (LAMs). One remarkable advancement in LAM is the second version of the deep potential with attention pre-trained model (DPA-2) developed by Zhang et al. from the DeepModeling open-source community [42, 43]. This model utilizes unified descriptors constructed by deep neural network architecture, and uses the multi-task pre-training strategy to jointly pre-train a multi-head model using multiple datasets that encompass multidisciplinary knowledge from a broad range of application domains. By fine-tuning for specific downstream tasks, the pre-trained DPA-2 model can be efficiently deployed to evaluate the potential energy surface (PES) of a specific research domain with precision and generalization.
In this work, we have developed an MLP based on the pre-trained DPA-2 model by fine-tuning methodology, named fine-tuned Fischer-Tropsch deep potential (FT
MATERIALS AND METHODS
DFT calculations
All spin-polarized DFT calculations were carried out by using Atomic-orbital Based Ab-initio Computation at UStc (ABACUS) package [44, 45]. The SG15-optimized Norm-Conserving Vanderbilt (SG15-ONCV) multi-projector pseudopotentials [46, 47] were employed and the valence configurations were [H]1s
DPA-2 and fine-tuning methodology
DPA-2 is a multi-task pre-trained LAM originating from the DP architecture and evolved from the DPA-1 model [52]. The DPA-1 descriptor has introduced an element-type embedding for encoding the elemental information covering the whole periodic table, and a gated self-attention mechanism [53] excelling in modeling the importance of neighboring atoms and re-weighting the interaction among them, which also makes the model generalizable and pre-trainable. Inheriting the DPA-1 backbone, the DPA-2 descriptor further enhances its resolution and generalizability of atomic representation through stacking multiple transformer [53] layers called representation transformer, incorporating operators such as convolution, symmetrization, localized self-attention, and gated self-attention, which can be interpreted as an E(3) equivariant graph neural network (GNN) and offers greater capacity compared to conventional GNNs [42], ensuring the robust capability of DPA-2 for serving as a LAM assembling comprehensive knowledge from massive pre-training data.
Besides having a more sophisticated model architecture, DPA-2 employs a multi-task training strategy for pre-training in multiple datasets labeled with different DFT settings to extract multidisciplinary knowledge. In particular, the multi-task DPA-2 model has multiple heads, and each head is an identical fitting network used to fit the DFT labels of each pre-training dataset from different downstream domains. During the pre-training process, the parameters within the DPA-2 descriptor are concurrently optimized through back-propagation using all pre-training datasets, while the parameters of the fitting network are updated exclusively with the specific pre-training dataset to which they are associated [42].
The pre-trained descriptor and fitting networks can be fine-tuned on specific downstream tasks, and the multidisciplinary knowledge learned from the multiple upstream datasets can help to reduce the consumption in model training and the amount of training data. In the fine-tuning process, the descriptor of the downstream model will be initialized with the pre-trained parameters, and the fitting network could also be initialized by choosing a fitting network in the pre-trained model. The energy bias in the fitting network will be aligned to the labels of the downstream dataset subsequently, and then the typical model training process will proceed with the initialized parameters incorporating upstream knowledge. Our FT
Double-to-single workflow for TS optimization
To investigate the catalytic reaction mechanism from a theoretical point of view, a crucial task is the optimization of TS structures of each elementary reaction for acquiring the activation energy. The climbing-image nudged elastic band (CI-NEB) [54, 55] and the dimer method [56, 57] are the two most popular TS optimization methods in heterogeneous catalysis simulation, representing two types of TS optimization methods respectively: double-ended methods that start from combining optimized initial state (IS) and final state (FS), and single-ended methods that are based on only one state, usually a guessed TS. It is well-known that the efficiency of single-ended methods highly relies on the quality of the input structure, but the optimization will converge quickly when the optimization reaches the quadratic region around TS. On the contrary, double-ended methods have no reliance on any guessed TS structure, but they tend to have convergence problems. In our work, we combine these two types of methods together as a workflow, named double-to-single (D2S), implemented in the atomic simulation environment (ASE) package [58] (as illustrated in Figure 1A) to make good use of them for accelerating the TS optimization process. The D2S workflow uses CI-NEB first to generate a rough reaction pathway with a relatively loose convergence criterion (usually 1.0 eV/Å for maximum of atomic forces), and then a single-ended method, such as the dimer, or a better choice, Sella algorithm based on iterative Hessian diagonalization and partitioned rational function optimization (P-RFO)[59, 60], is utilized by starting from the maximum point of the NEB pathway for strict TS optimization with target convergence criterion (usually 0.05 eV/Å for maximum of atomic forces). The free energy corrections, including zero-point energy (ZPE) and thermal corrections (translational, rotational and vibrational) for gas-phase molecules, as well as ZPE and vibrational contribution for adsorbates, are also computed using ASE. All these codes are open-sourced in the ATST-Tools suite [61], which supports using ABACUS and our FT
Figure 1. Scheme of model applications. (A) TS optimization, starting from a map of probable reaction network providing the reaction patterns of elementary reactions, and each reaction (taking CO dissociation as an example) is calculated by optimizing its IS, FS, and TS, where the TS is optimized by the D2S workflow; (B) Conventional and local schemes in optimizing the edge sites of the
Genetic algorithm for global optimization
We employed the genetic algorithm (GA) implemented in ASE [62], using FT
Ab initio atomistic thermodynamics
To evaluate the stability of reconstructed surfaces with different compositions, the ab initio atomistic thermodynamics theory developed by Reuter and Scheffler[66, 67] was used in this work. In this way, the surface energy (
where
For a surface that is in equilibrium with a bulk with a fixed composition (e.g., FeC
In our discussion, since not all structures contain the same amount of atoms, we defined a relative surface energy (
Here, the reference (
For convenience, we used the electronic energy of an isolated carbon atom (
RESULTS AND DISCUSSION
FT2DP construction and validation
Our model, FT
There are 30, 656 frames in our FT
A brief overview of the FT
Type of structures | 3D bulks | 2D surfaces | 1D strings | 0D clusters | All |
FT | |||||
Numbers of structures with Fe | 6, 917 | 13, 829 | 117 | 37 | 20, 900 |
Numbers of structures without Fe | 7, 341 | 1, 114 | 330 | 971 | 9, 756 |
Total numbers | 14, 258 | 14, 943 | 447 | 1, 008 | 30, 656 |
Figure 2. Sketch-map visualization of the FT
Our fine-tuning protocol was initialized by using the parameters of the global descriptor in pre-trained DPA-2.2.0 model and fitting network from the Domains_OC2M branch. The fine-tuning process on our dataset is done by following the default training process of the DPA-2 model with some setting modifications. In the default pre-training process of DPA-2.2.0 LAM, the learning rate starts from
To evaluate the generalizability of the FT
Figure 3. Validation results of the FT
Additionally, it is valuable to identify outlier structures with high prediction errors to investigate their characteristics and similarities. Structures in the entire FT
Validation results of the FT
Test items | FT | FT | Final FT |
FT | |||
Number of structures | 24, 525 | 6, 131 | 30, 656 |
Energy parity plot R | 1.000 | 1.000 | 1.000 |
Energy MAE (eV) | 0.1837 | 0.1890 | 0.1829 |
Energy RMSE (eV) | 0.3141 | 0.3233 | 0.3112 |
Energy MAE (meV/atom) | 5.497 | 5.780 | 5.500 |
Energy RMSE (meV/atom) | 10.02 | 10.85 | 9.974 |
Force parity plot R | 0.9605 | 0.9480 | 0.9570 |
Force MAE (eV/Å) | 0.0764 | 0.0792 | 0.0764 |
Force RMSE (eV/Å) | 0.1167 | 0.1271 | 0.1167 |
Reaction pathways
Many previous studies have demonstrated that certain surfaces exhibit much higher FTS activity, identifying them as the active surfaces[6, 71]. A well-known example is the
To further validate the accuracy of the FT
Figure 4. FTS reaction pathways on
Global optimization of reconstruction of edge sites
In addition to the
Before performing GA calculations, we first checked the accuracy of FT
There are two possible terminations of
Figure 5. Structure of the lowest-energy reconstruction of each surface and relative surface energies of low-energy reconstructions with respect to
Figure 6. Results of variable-composition calculations on the
The clean
Compared to other surfaces, the
CONCLUSIONS
To summarize, we have constructed the FT
We close the paper by giving a few general remarks. While the FT
DECLARATIONS
Acknowledgments
The authors acknowledge Dr. Duo Zhang from AI for Science Institute for helping with model training and dataset visualization and Dr. Yike Huang from AI for Science Institute for sharing template code on utilizing ASE-GA in surface systems. The authors also appreciate the support of the High-performance Computing Platform of Peking University and National Supercomputer Center in Tianjin for the computational resources.
Authors' contributions
Contributed equally to this work: Liu, Z.Q.; Deng, Z.
Dataset preparation, model training and evaluation, data analysis, writing: Liu, Z. Q.; Deng, Z.
Discussion of results, revision: Zhao, H.; Wang, H.; Chen, M.
Project conceptualization, methodology, supervision, revision, funding acquisition: Jiang, H.
Availability of data and materials
The dataset and model used in this study are both available on AIS Square (https://www.aissquare.com/). In detail, the dataset with the structural information, DFT label information, and DFT input setting files can be accessed at https://www.aissquare.com/datasets/detail?pageType=datasets&name=FT2DP-dataset-FeCHO-v1&id=306, and the model with its input scripts is available at https://www.aissquare.com/models/detail?pageType=models&id=307. Additionally, the codes for TS optimization are all available in the ATST-Tools repository (https://github.com/QuantumMisaka/ATST-Tools). All other data and codes supporting the findings presented in this work are available from the corresponding author upon reasonable request.
Financial support and sponsorship
This work is financially supported by the National Key Research and Development Program of China (Project no. 2022YFB4101401) and National Natural Science Foundation of China (Project no. 22273002).
Conflicts of interest
All authors declared that there are no conflicts of interest.
Ethical approval and consent to participate
Not applicable.
Consent for publication
Not applicable.
Copyright
® The Author(s) 2025.
Supplementary Materials
REFERENCES
1. Chen, B. W. J.; Xu, L.; Mavrikakis, M. Computational methods in heterogeneous catalysis. Chem. Rev. 2021, 121, 1007-48.
2. Mahmoudi, H.; Mahmoudi, M.; Doustdar, O.; et al. A review of Fischer Tropsch synthesis process, mechanism, surface chemistry and catalyst formulation. Biofuels. Eng. 2017, 2, 11-31.
3. Ail, S. S.; Dasappa, S. Biomass to liquid transportation fuel via Fischer Tropsch synthesis – technology review and current scenario. Renew. Sust. Energ. Rev. 2016, 58, 267-86.
4. Schulz, H. Comparing Fischer-Tropsch synthesis on iron- and cobalt catalysts: the dynamics of structure and function. Stud. Surf. Sci. Catal. 2007, 163, 177-99.
5. Ma, W.; Jacobs, G.; Sparks, D. E.; Todic, B.; Bukur, D. B.; Davis, B. H. Quantitative comparison of iron and cobalt based catalysts for the Fischer-Tropsch synthesis under clean and poisoning conditions. Catal. Today. 2020, 343, 125-36.
6. Liu, Q. Y.; Shang, C.; Liu, Z. P. In. situ. active. site. for. Fe-catalyzed. Fischer–Tropsch. synthesis:. recent. progress. and. future. challenges. J. Phys. Chem. Lett. 2022, 13, 3342-52.
7. de Smit, E.; Cinquini, F.; Beale, A. M.; et al. Stability and reactivity of
8. Zhang, J.; Abbas, M.; Zhao, W.; Chen, J. Enhanced stability of a fused iron catalyst under realistic Fischer–Tropsch synthesis conditions: insights into the role of iron phases (
9. Yang, C.; Zhao, H.; Hou, Y.; Ma, D. Fe
10. Zhao, H.; Liu, J. X.; Yang, C.; et al. Synthesis of iron-carbide nanoparticles: identification of the active phase and mechanism of Fe-based Fischer–Tropsch synthesis. CCS. Chem. 2021, 3, 2712-24.
11. Cano, L. A.; Cagnoli, M. V.; Fellenz, N. A.; et al. Influence of the crystal size of iron active species on the activity and selectivity. Appl. Catal. A. Gen. 2010, 379, 105-10.
12. Torres Galvis, H. M.; Bitter, J. H.; Davidian, T.; Ruitenbeek, M.; Dugulan, A. I.; de Jong, K. P. Iron particle size effects for direct production of lower olefins from synthesis gas. J. Am. Chem. Soc. 2012, 134, 16207-15.
13. Park, J. C.; Yeo, S. C.; Chun, D. H.; et al. Highly activated K-doped iron carbide nanocatalysts designed by computational simulation for Fischer–Tropsch synthesis. J. Mater. Chem. A. 2014, 2, 14371-9.
14. Pham, T. H.; Qi, Y.; Yang, J.; et al. Insights into Hägg iron-carbide-catalyzed Fischer–Tropsch synthesis: suppression of CH
15. Song, N.; Cao, J.; Chen, B.; Qian, G.; Duan, X.; Zhou, X. CO adsorption and activation of
16. Chen, B.; Wang, D.; Duan, X.; et al. Charge-tuned CO activation over a
17. Li, T.; Wen, X.; Yang, Y.; Li, Y. W.; Jiao, H. Mechanistic aspects of CO activation and C–C bond formation on the Fe/C- and Fe-terminated Fe
18. Yin, J.; Liu, X.; Liu, X. W.; et al. Theoretical exploration of intrinsic facet-dependent CH
19. Zhang, X.; Wang, L.; Helwig, J.; et al. Artificial intelligence for science in quantum, atomistic, and continuum systems. arXiv2023, arXiv: 2307.08423. Available online: https://arxiv.org/abs/2307.08423. (accessed on 17 Mar 2025).
20. Ma, S.; Liu, Z. P. Machine learning for atomic simulation and activity prediction in heterogeneous catalysis: current status and future. ACS. Catal. 2020, 10, 13213-26.
21. Huang, S. D.; Shang, C.; Kang, P. L.; Zhang, X. J.; Liu, Z. P. LASP: fast global potential energy surface exploration. WIREs. Comput. Mol. Sci. 2019, 9, e1415.
22. Huang, S. D.; Shang, C.; Kang, P. L.; Liu, Z. P. Atomic structure of boron resolved using machine learning and global sampling. Chem. Sci. 2018, 9, 8644-55.
23. Chen, D.; Chen, L.; Zhao, Q. C.; Yang, Z. X.; Shang, C.; Liu, Z. P. Square-pyramidal subsurface oxygen[Ag
24. Liu, Q. Y.; Shang, C.; Liu, Z. P. In situ active site for CO activation in Fe-catalyzed Fischer-Tropsch synthesis from machine learning. J. Am. Chem. Soc. 2021, 143, 11109-20.
25. Liu, Q. Y.; Chen, D.; Shang, C.; Liu, Z. P. An optimal Fe–C coordination ensemble for hydrocarbon chain growth: a full Fischer–Tropsch synthesis mechanism from machine learning. Chem. Sci. 2023, 14, 9461-75.
26. van Steen, E.; Schulz, H. Polymerisation kinetics of the Fischer-Tropsch CO hydrogenation using iron and cobalt based catalysts. Appl. Catal. A. Gen. 1999, 186, 309-20.
27. Han, J.; Zhang, L.; Car, R.; Weinan, E. Deep potential: a general representation of a many-body potential energy surface. Commun. Comput. Phys. 2018, 23, 629-39.
28. Wang, H.; Zhang, L.; Han, J.; Weinan, E. DeePMD-kit: a deep learning package for many-body potential energy representation and molecular dynamics. Comput. Phys. Commun. 2018, 228, 178-84.
29. Zeng, J.; Zhang, D.; Lu, D.; et al. DeePMD-kit v2: a software package for deep potential models. J. Chem. Phys. 2023, 159, 054801.
30. DeePMD-kit. https://github.com/deepmodeling/deepmd-kit. (accessed on 2025-03-17).
31. Zhang, L.; Han, J.; Wang, H.; Saidi, W. A.; Car, R.; Weinan, E. End-to-end symmetry preserving inter-atomic potential energy model for finite and extended systems. arXiv2018, arXiv: 1805.09003. Available online: https://arxiv.org/abs/1805.09003. (accessed on 17 Mar 2025).
32. Hedman, D.; McLean, B.; Bichara, C.; Maruyama, S.; Larsson, J. A.; Ding, F. Dynamics of growing carbon nanotube interfaces probed by machine learning-enabled molecular simulations. Nat. Commun. 2024, 15, 4076.
33. Liu, J. C.; Luo, L.; Xiao, H.; Zhu, J.; He, Y.; Li, J. Metal affinity of support dictates sintering of gold catalysts. J. Am. Chem. Soc. 2022, 144, 20601-9.
34. Zhou, L.; Fu, X. P.; Wang, R.; et al. Dynamic phase transitions dictate the size effect and activity of supported gold catalysts. Sci. Adv. 2024, 10, eadr4145.
35. Hou, P.; Yu, Q.; Luo, F.; Liu, J. C. Reactant-induced dynamic active sites on Cu catalysts during the water - gas shift reaction. ACS. Catal. 2025, 15, 352-60.
36. Wu, J.; Chen, D.; Chen, J.; Wang, H. Structural and composition evolution of palladium catalyst for CO oxidation under steady-state reaction conditions. J. Phys. Chem. C. 2023, 127, 6262-70.
37. Zhang, Y.; Wang, H.; Chen, W.; et al. DP-GEN: a concurrent learning platform for the generation of reliable deep learning based potential energy models. Comput. Phys. Commun. 2020, 253, 107206.
38. Cheng, X.; Wu, C.; Xu, J.; Han, Y.; Xie, W.; Hu, P. Leveraging machine learning potentials for in-situ searching of active sites in heterogeneous catalysis. Precis. Chem. 2024, 2, 570-86.
39. Chanussot, L.; Das, A.; Goyal, S.; et al. Open Catalyst 2020 (OC20) dataset and community challenges. ACS. Catal. 2021, 11, 6059-72.
40. Tran, R.; Lan, J.; Shuaibi, M.; et al. The Open Catalyst 2022 (OC22) dataset and challenges for oxide electrocatalysts. ACS. Catal. 2023, 13, 3066-84.
41. Deng, B.; Zhong, P.; Jun, K.; et al. CHGNet as a pretrained universal neural network potential for charge-informed atomistic modelling. Nat. Mach. Intell. 2023, 5, 1031-41.
42. Zhang, D.; Liu, X.; Zhang, X.; et al. DPA-2: a large atomic model as a multi-task learner. npj. Comput. Mater. 2024, 10, 293.
43. DeepModeling Community. https://deepmodeling.com. (accessed on 2025-03-17).
44. Chen, M.; Guo, G. C.; He, L. Systematically improvable optimized atomic basis sets for ab initio calculations. J. Phys. Condens. Mat. 2010, 22, 445501.
45. Li, P.; Liu, X.; Chen, M.; et al. Large-scale ab initio simulations based on systematically improvable atomic basis. Comput. Mater. Sci. 2016, 112, 503-17.
46. Schlipf, M.; Gygi, F. Optimization algorithm for the generation of ONCV pseudopotentials. Comput. Phys. Commun. 2015, 196, 36-44.
47. SG15 ONCV Potentials. http://quantum-simulation.org/potentials/sg15_oncv. (accessed on 2025-03-17).
48. Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865.
49. Lin, P.; Ren, X.; He, L. Strategy for constructing compact numerical atomic orbital basis sets by incorporating the gradients of reference wavefunctions. Phys. Rev. B. 2021, 103, 235131.
50. Monkhorst, H. J.; Pack, J. D. Special points for Brillouin-zone integrations. Phys. Rev. B. 1976, 13, 5188.
51. Methfessel, M.; Paxton, A. T. High-precision sampling for Brillouin-zone integration in metals. Phys. Rev. B. 1989, 40, 3616.
52. Zhang, D.; Bi, H.; Dai, F. Z.; et al. Pretraining of attention-based deep learning potential model for molecular simulation. npj. Comput. Mater. 2024, 10, 94.
53. Vaswani, A.; Shazeer, N.; Parmar, N.; et al. Attention is all you need. arXiv2017, arXiv: 1706.03762. Available online: https://arxiv.org/abs/1706.03762. (accessed on 17 Mar 2025).
54. Henkelman, G.; Uberuaga, B. P.; Jónsson, H. A climbing image nudged elastic band method for finding saddle points and minimum energy paths. J. Chem. Phys. 2000, 113, 9901-4.
55. Henkelman, G.; Jónsson, H. Improved tangent estimate in the nudged elastic band method for finding minimum energy paths and saddle points. J. Chem. Phys. 2000, 113, 9978-85.
56. Henkelman, G.; Jónsson, H. A dimer method for finding saddle points on high dimensional potential surfaces using only first derivatives. J. Chem. Phys. 1999, 111, 7010-22.
57. Kästner, J.; Sherwood, P. Superlinearly converging dimer method for transition state search. J. Chem. Phys. 2008, 128, 014106.
58. Larsen, A. H.; Mortensen, J. J.; Blomqvist, J.; et al. The atomic simulation environment - a Python library for working with atoms. J. Phys. Condens. Mat. 2017, 29, 273002.
59. Hermes, E. D.; Sargsyan, K.; Najm, H. N.; Zádor, J. Accelerated saddle point refinement through full exploitation of partial hessian diagonalization. J. Chem. Theory. Comput. 2019, 15, 6536-49.
60. Hermes, E. D.; Sargsyan, K.; Najm, H. N.; Zádor, J. Sella, an open-source automation-friendly molecular saddle point optimizer. J. Chem. Theory. Comput. 2022, 18, 6974-88.
61. ATST-Tools. https://github.com/QuantumMisaka/ATST-Tools. (accessed on 2025-03-17).
62. Vilhelmsen, L. B.; Hammer, B. A genetic algorithm for first principles global structure optimization of supported nano structures. J. Chem. Phys. 2014, 141, 044711.
63. Zakaryan, H. A.; Kvashnin, A. G.; Oganov, A. R. Stable reconstruction of the (110) Surface and its role in pseudocapacitance of rutile-like RuO
64. Hinuma, Y.; Kamachi, T.; Hamamoto, N.; Takao, M.; Toyao, T.; Shimizu, K. Surface oxygen vacancy formation energy calculations in 34 orientations of
65. Momma, K.; Izumi, F. VESTA 3 for three-dimensional visualization of crystal, volumetric and morphology data. J. Appl. Cryst. 2011, 44, 1272-6.
66. Reuter, K.; Scheffler, M. Composition, structure, and stability of RuO
67. Reuter, K.; Scheffler, M. Composition and structure of the RuO
68. AIS-Square. DPA-2.2.0 pre-trained model. https://www.aissquare.com/models/detail?pageType=models&name=DPA-2.2.0-v3.0.0b3&id=272. (accessed on 2025-03-17).
69. AIS-Square. OpenLAM Project. https://www.aissquare.com/openlam. (accessed on 2025-03-17).
70. Kulichenko, M.; Nebgen, B.; Lubbers, N.; et al. Data generation for machine learning interatomic potentials and beyond. Chem. Rev. 2024, 124, 13681-714.
71. Broos, R. J. P.; Zijlstra, B.; Filot, I. A. W.; Hensen, E. J. M. Quantum-chemical DFT study of direct and H- and C-assisted CO dissociation on the
72. Zhao, Z. J.; Li, Z.; Cui, Y.; et al. Importance of metal-oxide interfaces in heterogeneous catalysis: a combined DFT, microkinetic, and experimental study of water-gas shift on Au/MgO. J. Catal. 2017, 345, 157-69.
73. Chen, C.; Zuo, Y.; Ye, W.; Li, X.; Ong, S. P. Learning properties of ordered and disordered materials from multi-fidelity data. Nat. Comp. Sci. 2021, 1, 46-53.
74. Xie, J.; Yang, J.; Dugulan, A. I.; et al. Size and promoter effects in supported iron Fischer-Tropsch catalysts: insights from experiment and theory. ACS. Catal. 2016, 6, 3147-57.
75. Zhao, S.; Liu, X. W.; Huo, C. F.; Li, Y. W.; Wang, J.; Jiao, H. Surface morphology of Hägg iron carbide (
76. Morgenstern, K.; Rosenfeld, G.; Comsa, G. Decay of two-dimensional Ag islands on Ag(111). Phys. Rev. Lett. 1996, 76, 2113.
Cite This Article
How to Cite
Liu, Z. Q.; Deng, Z.; Zhao, H.; Wang, H.; Chen, M.; Jiang, H. FT2DP: large atomic model fine-tuned machine learning potential for accelerating atomistic simulation of iron-based Fischer-Tropsch synthesis. J. Mater. Inf. 2025, 5, 27. http://dx.doi.org/10.20517/jmi.2024.105
Download Citation
Export Citation File:
Type of Import
Tips on Downloading Citation
Citation Manager File Format
Type of Import
Direct Import: When the Direct Import option is selected (the default state), a dialogue box will give you the option to Save or Open the downloaded citation data. Choosing Open will either launch your citation manager or give you a choice of applications with which to use the metadata. The Save option saves the file locally for later use.
Indirect Import: When the Indirect Import option is selected, the metadata is displayed and may be copied and pasted as needed.
About This Article
Special Issue
Copyright
Data & Comments
Data

Comments
Comments must be written in English. Spam, offensive content, impersonation, and private information will not be permitted. If any comment is reported and identified as inappropriate content by OAE staff, the comment will be removed without notice. If you have any queries or need any help, please contact us at [email protected].