research papers\(\def\hfill{\hskip 5em}\def\hfil{\hskip 3em}\def\eqno#1{\hfil {#1}}\)

IUCrJ
Volume 11| Part 2| March 2024| Pages 182-189
ISSN: 2052-2525

The prediction of single-molecule magnet properties via deep learning

crossmark logo

aDepartment of Chemistry, Tokyo University of Science, 1-3 Kagurazaka, Shinjuku-ku, Tokyo 1628601, Japan
*Correspondence e-mail: akitsu2@rs.tus.ac.jp

Edited by C.-Y. Su, Sun Yat-Sen University, China (Received 12 December 2023; accepted 22 January 2024; online 1 February 2024)

This paper uses deep learning to present a proof-of-concept for data-driven chemistry in single-molecule magnets (SMMs). Previous discussions within SMM research have proposed links between molecular structures (crystal structures) and single-molecule magnetic properties; however, these have only interpreted the results. Therefore, this study introduces a data-driven approach to predict the properties of SMM structures using deep learning. The deep-learning model learns the structural features of the SMM molecules by extracting the single-molecule magnetic properties from the 3D coordinates presented in this paper. The model accurately determined whether a molecule was a single-molecule magnet, with an accuracy rate of approximately 70% in predicting the SMM properties. The deep-learning model found SMMs from 20 000 metal complexes extracted from the Cambridge Structural Database. Using deep-learning models for predicting SMM properties and guiding the design of novel molecules is promising.

1. Introduction

Single-molecule magnets (SMMs) are metal complexes that exhibit magnetic relaxation behaviour like bulk ferromagnets as single molecules (Gatteschi et al., 2006[Gatteschi, D., Sessoli, R. & Villain, J. (2006). Molecular Nanomagnets, Oxford University Press.]; Winpenny & Aromí, 2006[Winpenny, R. & Aromí, G. (2006). Single-Molecule Magnets and Related Phenomena, p. 122. Berlin: Springer.]). Molecular-sized nanoscale magnets are expected to find applications in high-density memory and quantum-computing devices (Leuenberger & Loss, 2001[Leuenberger, M. N. & Loss, D. (2001). Nature, 410, 789-793.]; Eliseeva & Bünzli, 2011[Eliseeva, S. V. & Bünzli, J. G. (2011). New J. Chem. 35, 1165-1176.]). The effective activation energy barrier (Ueff) is typically used to assess magnetization reversal in SMMs (Woodruff et al., 2013[Woodruff, D. N., Winpenny, R. E. & Layfield, R. A. (2013). Chem. Rev. 113, 5110-5148.]). Owing to the relationship shown in equation (1)[link] between Ueff, zero-field splitting parameter D and total spin quantum number S of the molecule (Zhang et al., 2013[Zhang, P., Guo, Y.-N. & Tang, J. (2013). Coord. Chem. Rev. 257, 1728-1763.]), D and S are essential keywords in SMM molecule designs:

[U = \left| D \right|\,{S^2}. \eqno (1)]

Although early strategies concentrated on multinuclear complexes of transition metals, such as those represented by the Mn12 cluster (Sessoli et al., 1993[Sessoli, R., Gatteschi, D., Caneschi, A. & Novak, M. A. (1993). Nature, 365, 141-143.]), mononuclear lanthanide complexes now play a central role (Zhang et al., 2013[Zhang, P., Guo, Y.-N. & Tang, J. (2013). Coord. Chem. Rev. 257, 1728-1763.]; Dey et al., 2018[Dey, A., Kalita, P. & Chandrasekhar, V. (2018). ACS Omega, 3, 9462-9475.]). Ab initio calculations have demonstrated that highly axial [DyO]+ ions in lanthanide metal complexes display Ueff values above 3000 K (Ungur & Chibotaru, 2011[Ungur, L. & Chibotaru, L. F. (2011). Phys. Chem. Chem. Phys. 13, 20086-20090.]). Molecules with Ueff values of approximately 2000 K have been reported (Randall McClain et al., 2018[Randall McClain, K. R., Gould, C. A., Chakarawet, K., Teat, S. J., Groshens, T. J., Long, J. R. & Harvey, B. G. (2018). Chem. Sci. 9, 8492-8503.]). However, it is exceedingly challenging to fabricate molecules that exhibit high Ueff values because the Ueff of traditional SMMs is typically in the range of tens to hundreds of kelvin (Tang & Zhang, 2015[Tang, J. & Zhang, P. (2015). Lanthanide Single Molecule Magnets, 2 Berlin: Springer.]; Liu et al., 2015[Liu, K., Shi, W. & Cheng, P. (2015). Coord. Chem. Rev. 289-290, 74-122.]; Parmar et al., 2021[Parmar, V. S., Mills, D. P. & Winpenny, R. E. P. (2021). Chem. A Eur. J. 27, 7625-7645.]). Ueff is affected by quantum tunnelling and antiferromagnetic interactions, and achieving control over the structure of lanthanide metal complexes in crystals is challenging. Currently, there are no established guidelines for SMM design.

In the theoretical field, discussions using magnetic properties obtained from the complete active space self-consistent field (CASSCF) method (Roos & Malmqvist, 2004[Roos, B. O.-Å. & Malmqvist, P. (2004). Phys. Chem. Chem. Phys. 6, 2919-2927.]) using the Molcas program (Aquilante et al., 2010[Aquilante, F., De Vico, L., Ferré, N., Ghigo, G., Malmqvist, P., Neogrády, P., Pedersen, T. B., Pitoňák, M., Reiher, M., Roos, B. O., Serrano-Andrés, L., Urban, M., Veryazov, V. & Lindh, R. (2010). J. Comput. Chem. 31, 224-247.]) provide insight into the origin of SMMs. In addition, some techniques leverage the CASSCF method to estimate Ueff directly (Aravena, 2018[Aravena, D. (2018). J. Phys. Chem. Lett. 9, 5327-5333.]; Yin & Li, 2020[Yin, B. & Li, C. C. (2020). Phys. Chem. Chem. Phys. 22, 9923-9933.]). Nevertheless, these methods require substantial computational resources and time, and selecting computational conditions is a significant bottleneck. These algorithms are unsuitable for extensive calculations that require a vast search space, such as molecular design, because they have only been utilized to interpret experimental outcomes.

Recently, interest in data-driven chemistry and theory-predicate research, including machine learning, has grown (Engel, 2006[Engel, T. (2006). J. Chem. Inf. Model. 46, 2267-2277.]; Mitchell, 2014[Mitchell, J. B. (2014). WIREs Comput. Mol. Sci. 4, 468-481.]). The Chemical Reviews journal released a special issue on `Machine Learning at the Atomic Scale' (Ceriotti et al., 2021[Ceriotti, M., Clementi, C. & Anatole von Lilienfeld, O. (2021). Chem. Rev. 121, 9719-9721.]). The Tsuda group at RIKEN provided instances of newly designed fluorescent materials successfully developed using AI and density functional theory (Sumita et al., 2018[Sumita, M., Yang, X., Ishihara, S., Tamura, R. & Tsuda, K. (2018). ACS Cent. Sci. 4, 1126-1133.]; Sumita et al., 2022[Sumita, M., Terayama, K., Suzuki, N., Ishihara, S., Tamura, R., Chahal, M. K., Payne, D. T., Yoshizoe, K. & Tsuda, K. (2022). Sci. Adv. 8, eabj3906.]). However, data-driven research at the laboratory level has remained unsuccessful. Machine learning of metal complexes has predominantly focused on investigating hexacoordinated and tetracoordinated complexes (Nandy et al., 2021[Nandy, A., Duan, C., Taylor, M. G., Liu, F., Steeves, A. H. & Kulik, H. J. (2021). Chem. Rev. 121, 9927-10000.]). However, addressing the structural changes caused by the spin multiplicity and similarity definitions is crucial for effectively performing machine learning on these complexes. Thus, extending conventional organic molecule machine learning to all elements is insufficient (Janet et al., 2019[Janet, J. P., Liu, F., Nandy, A., Duan, C., Yang, T., Lin, S. & Kulik, H. J. (2019). Inorg. Chem. 58, 10592-10606.]). It is unclear what AI can do in the SMM field, in which lanthanide ions and metal complexes with complex structures play central roles.

Here, we concentrate on the SMM research and present a proof-of-concept for the data-driven chemistry of lanthanide metal complexes and metal complexes with intricate structures. A dataset was compiled from papers on salen-type SMMs from the preceding decade, and deep learning was used to predict SMM behaviour. The SMM characteristics were predicted using approximately 20 000 crystal structures of the Schiff base metal complexes extracted from the Cambridge Structural Database (CSD). Although many statistical investigations (Duan et al., 2022[Duan, Y., Rosaleny, L. E., Coutinho, J. T., Giménez-Santamarina, S., Scheie, A., Baldoví, J. J., Cardona-Serra, S. & Gaita-Ariño, A. (2022). Nat. Commun. 13, 7626.]) and systematic reviews (Zhang et al., 2013[Zhang, P., Guo, Y.-N. & Tang, J. (2013). Coord. Chem. Rev. 257, 1728-1763.]; Dey et al., 2018[Dey, A., Kalita, P. & Chandrasekhar, V. (2018). ACS Omega, 3, 9462-9475.]) of SMMs have been conducted, this is the first report of deep learning on the molecular structures of SMMs. From the CSD, numerous molecules predicted to display SMM behaviour did indeed exhibit SMM behaviour, thereby demonstrating the capacity of the model to understand SMM features and `discover' SMMs. Such indications imply that implementing AI in SMM research could introduce a novel paradigm shift akin to AlphaGo's AI (Silver et al., 2016[Silver, D., Huang, A., Maddison, C. J., Guez, A., Sifre, L., van den Driessche, G., Schrittwieser, J., Antonoglou, I., Panneershelvam, V., Lanctot, M., Dieleman, S., Grewe, D., Nham, J., Kalchbrenner, N., Sutskever, I., Lillicrap, T., Leach, M., Kavukcuoglu, K., Graepel, T. & Hassabis, D. (2016). Nature, 529, 484-489.]) that surpassed human players.

2. Methods

2.1. Datasets

Molecules reported as SMMs comprise numerous chemical species like phthalocyanine complexes, metallocenes and polyoxometalate complexes (Duan et al., 2022[Duan, Y., Rosaleny, L. E., Coutinho, J. T., Giménez-Santamarina, S., Scheie, A., Baldoví, J. J., Cardona-Serra, S. & Gaita-Ariño, A. (2022). Nat. Commun. 13, 7626.]). Salen-type metal complexes were selected as the focus of this study. Salen-type ligands can be effortlessly synthesized from aldehydes and amines to form complexes with various 3d and 4f metals. Several SMM structures have been reported, including mononuclear, binuclear and 3d–4f multinuclear complexes (Liu et al., 2015[Liu, K., Shi, W. & Cheng, P. (2015). Coord. Chem. Rev. 289-290, 74-122.]). From these varied structures, it is expected that structures specific to SMMs can be identified.

The dataset was created using approximately 800 papers from 2011–2021 that were found using the keywords `salen + SMM' using Google Scholar. Data on the crystal structures and whether they exhibited SMM behaviour were collected from these studies, and a dataset was created. Non-SMM molecules were defined as those that did not exhibit magnetic relaxation behaviour according to AC magnetic susceptibility measurements. Tunnelling effects are suppressed under an appropriate DC magnetic field, increasing the likelihood of SMM behaviour. If the measurement results exist under both zero and DC magnetic fields, priority is given to the results under the zero magnetic field.

CIF files were obtained from the CSD (Groom et al., 2016[Groom, C. R., Bruno, I. J., Lightfoot, M. P. & Ward, S. C. (2016). Acta Cryst. B72, 171-179.]). The crystal structures were verified using the Mercury crystal drawing software (Macrae et al., 2020[Macrae, C. F., Sovago, I., Cottrell, S. J., Galek, P. T. A., McCabe, P., Pidcock, E., Platings, M., Shields, G. P., Stevens, J. S., Towler, M. & Wood, P. A. (2020). J. Appl. Cryst. 53, 226-235.]). The molecular structures were then converted to molecular structure files (XYZ files) by excluding non-pertinent molecules such as crystal water and were utilized as coordinate data. The explanatory variables were the molecular structures retrieved from the crystal structures, while the objective variable was the SMM behaviour.

2.2. Input representation

The input representations of the molecular structures typically include molecular descriptors, graphs and SMILES (Elton et al., 2019[Elton, D. C., Boukouvalas, Z., Fuge, M. D. & Chung, P. W. (2019). Mol. Syst. Des. Eng. 4, 828-849.]). This study utilized 3D images (voxels) as input representations to represent the metal complexes. Although their memory-intensive requirements compared with other input representations, voxel usage is expected to retain the 3D information of the molecules. Their use has been demonstrated for compounds of different sizes and structures, including organic molecules, inorganic solids and proteins (Ryan et al., 2018[Ryan, K., Lengyel, J. & Shatruk, M. (2018). J. Am. Chem. Soc. 140, 10158-10168.]; Amidi et al., 2018[Amidi, A., Amidi, S., Vlachakis, D., Megalooikonomou, V., Paragios, N. & Zacharaki, E. I. (2018). PeerJ, 6, e4750.]; Kuzminykh et al., 2018[Kuzminykh, D., Polykovskiy, D., Kadurin, A., Zhebrak, A., Baskov, I., Nikolenko, S., Shayakhmetov, R. & Zhavoronkov, A. (2018). Mol. Pharm. 15, 4378-4385.]; Li et al., 2021[Li, C., Wang, J., Niu, Z., Yao, J. & Zeng, X. (2021). Brief. Bioinform. 22, bbab078.]; Park & Seok, 2022[Park, S. & Seok, C. (2022). J. Chem. Inf. Model. 62, 3157-3168.]). These results represent the structural features of metal complexes with different central metals and structures.

The dimensions of the voxels were set to 64 × 64 × 64, with the element type represented by the assigned colour (RGB value) for each element. One side of the voxel was established at 12 Å to preserve the molecular structure. 3D molecular images were created by placing spherical voxels at 3D atomic coordinates obtained from the XYZ file, and 3D molecular images were created (Fig. 1[link]). A balance between voxel resolution and computational resources must be achieved, and an appropriate voxel size must be determined via trial-and-error experimentation.

[Figure 1]
Figure 1
Voxel diagram of the metal complex. Create a 64 × 64 × 64 × 3 array and specify the colour of each element [R, G, B] for the voxels within a radius r centred on the atom position. [0,0,0] is specified for empty space. In addition, the size of the sphere was changed for each element in proportion to the ionic radius. It is an intuitive and highly versatile descriptor, requiring no special pre-processing.

2.3. Model architecture

A 3D convolutional neural network (3D-CNN) was used as the deep-learning model to create a binary classification model that predicts whether a molecule is an SMM based on its molecular structure (Fig. 2[link]); this multilayer neural network consists of convolutional layers and fully connected layers. The input image is output as a scalar value between 0 and 1 by the last fully connected layer after the 3D convolutional layer extracts the 3D image features (Jogin et al., 2018[Jogin, M., Madhulika, M., Divya, G., Meghana, R. & Apoorva, S. (2018). Proceedings of the 3rd IEEE International Conference on Recent Trends in Electronics, Information & Communication Technology (RTEICT), 18-19 May 2018, Bengaluru, India, pp. 2319-2323. IEEE Press.]). In this study, SMMs were set to 0 and non-SMMs to 1; however, which was 0 did not affect the results. ResNet with residue connections (He et al., 2019[He, K., Zhang, X., Ren, S. & Sun, J. (2015). Proceedings of the IEEE conference on computer vision and Pattern Recognition (CVPR2015), 7-12 June 2024, Boston, MA, USA, pp. 70-778. IEEE Press.]) was selected as the network structure owing to its impressive image classification accuracy. ResNet can effectively construct deep CNN models with many layers through residual connections between the convolutional layers. The model comprises multiple Res-blocks connected via the GAP layer to fully connected layers. The Res-blocks employ the activation function pre-positioning type, with the ReLU layer positioned before the confluence. Furthermore, a bottleneck-type structure was adopted for these blocks to reduce computational demands. Batch normalization (Bjorck et al., 2018[Bjorck, N., Gomes, C. P., Selman, B. & Weinberger, K. Q. (2018). Advances in Neural Information Processing Systems 31 (NeurIPS 2018), 3-8 December 2018, Montreal, Canada, pp. 31.]) and a 20% dropout (Gal & Ghahramani, 2015[Gal, Y. & Ghahramani, Z. (2015). arXiv: 1506.02158.]) were used to prevent overlearning. The ReLU function (Agarap, 2019[Agarap, A. F. (2019). Deep learning using rectified linear units (ReLU), arXiv: 1803.08375.]) was used as the activation function, and the sigmoid function (Cybenko, 1989[Cybenko, G. (1989). Math. Contr. Signal. Syst. 2, 303-314.]) was used as the final layer.

[Figure 2]
Figure 2
Conceptual diagram of SMM prediction. The CIF file was edited using crystal drawing software such as Mercury. Crystal water contained in the crystal was removed, and the structure of only the target complex molecule was converted into an XYZ file. The XYZ file is converted into voxels, and 3D-CNN predicts whether it is SMM or non-SMM.

2.4. Model training

As the dataset used in this study was unbalanced data with SMM:non-SMM of 2:1, undersampling (Mohammed et al., 2020[Mohammed, R., Rawashdeh, J. & Abdullah, M. (2020). Proceedings of the 11th International Conference on Information and Communications Systems (ICICS2020), 7-9 April 2020, Irbid, Jordan, pp. 243-248. IEEE Press.]) was used to reduce the SMM data to eliminate data unbalances. The dataset was split 6:2:2 into training, validation and test data. The CNN was trained with the Adam optimiser using AMSGrad (Reddi et al., 2019[Reddi, S. J., Kale, S. & Kumar, S. (2019). arXiv: 1904.09237.]) with hyperparameters ɛ = 1 × 10−7, β1 = 0.9 and β2 = 0.999. The learning rate was reduced from an initial value of 1 × 10−2 to 1 × 10−5 using the cosine reduction rate (Loshchilov & Hutter, 2016[Loshchilov, I. & Hutter, F. (2016). arXiv: 1608.03983.]). The batch size was set to eight to account for GPU memory. The loss function used was the cross-entropy error (Bishop, 1995[Bishop, C. M. (1995). Neural Networks for Pattern Recognition. Oxford University Press.]), which was trained to minimize. The correct response rate and AUC were used as model evaluation indices. Data augmentation (Shorten & Khoshgoftaar, 2019[Shorten, C. & Khoshgoftaar, T. M. (2019). J. Big Data, 6, 60.]) was performed during training to reduce over-learning. As the molecular structure was obtained in atomic coordinates, voxels were generated after rotation, scaling and translation to 3D coordinates. By performing data augmentation, the model is expected to prevent overlearning and acquire rotational and translational invariance. The CNN was trained for 1000 epochs. These hyperparameters were derived by trial and error.

2.5. Software and libraries

All the programs were implemented using Python 3. TensorFlow 2.0 (Abadi et al., 2016[Abadi, M., Barham, P., Chen, J., Chen, Z., Davis, A., Dean, J., Devin, M., Ghemawat, S., Irving, G. & Isard, M. (2016). Proceedings of the 12th USENIX Symposium on Operating Systems Design and Implementation (OSDI16), 2-4 November 2016, Savannah, GA, USA, pp. 265-283. The USENIX Association.]) was used as the machine learning library. The NumPy (Harris et al., 2020[Harris, C. R., Millman, K. J., van der Walt, S. J., Gommers, R., Virtanen, P., Cournapeau, D., Wieser, E., Taylor, J., Berg, S., Smith, N. J., Kern, R., Picus, M., Hoyer, S., van Kerkwijk, M. H., Brett, M., Haldane, A., del Río, J. F., Wiebe, M., Peterson, P., Gérard-Marchant, P., Sheppard, K., Reddy, T., Weckesser, W., Abbasi, H., Gohlke, C. & Oliphant, T. E. (2020). Nature, 585, 357-362.]) and SciPy (Virtanen et al., 2020[Virtanen, P., Gommers, R., Oliphant, T. E., Haberland, M., Reddy, T., Cournapeau, D., Burovski, E., Peterson, P., Weckesser, W., Bright, J., van der Walt, S. J., Brett, M., Wilson, J., Millman, K. J., Mayorov, N., Nelson, A. R. J., Jones, E., Kern, R., Larson, E., Carey, C. J., Polat, İ., Feng, Y., Moore, E. W., VanderPlas, J., Laxalde, D., Perktold, J., Cimrman, R., Henriksen, I., Quintero, E. A., Harris, C. R., Archibald, A. M., Ribeiro, A. H., Pedregosa, F., van Mulbregt, P., Vijaykumar, A., Bardelli, A. P., Rothberg, A., Hilboll, A., Kloeckner, A., Scopatz, A., Lee, A., Rokem, A., Woods, C. N., Fulton, C., Masson, C., Häggström, C., Fitzgerald, C., Nicholson, D. A., Hagen, D. R., Pasechnik, D. V., Olivetti, E., Martin, E., Wieser, E., Silva, F., Lenders, F., Wilhelm, F., Young, G., Price, G. A., Ingold, G., Allen, G. E., Lee, G. R., Audren, H., Probst, I., Dietrich, J. P., Silterra, J., Webber, J. T., Slavič, J., Nothman, J., Buchner, J., Kulick, J., Schönberger, J. L., de Miranda Cardoso, J. V., Reimer, J., Harrington, J., Rodríguez, J. L. C., Nunez-Iglesias, J., Kuczynski, J., Tritz, K., Thoma, M., Newville, M., Kümmerer, M., Bolingbroke, M., Tartre, M., Pak, M., Smith, N. J., Nowaczyk, N., Shebanov, N., Pavlyk, O., Brodtkorb, P. A., Lee, P., McGibbon, R. T., Feldbauer, R., Lewis, S., Tygier, S., Sievert, S., Vigna, S., Peterson, S., More, S., Pudlik, T., Oshima, T., Pingel, T. J., Robitaille, T. P., Spura, T., Jones, T. R., Cera, T., Leslie, T., Zito, T., Krauss, T., Upadhyay, U., Halchenko, Y. O. & Vázquez-Baeza, Y. (2020). Nat. Methods, 17, 261-272.]) libraries were used for coordinate transformation and voxel generation. Open Babel (O'Boyle et al., 2011[O'Boyle, N. M., Banck, M., James, C. A., Morley, C., Vandermeersch, T. & Hutchison, G. R. (2011). J. Cheminform, 3, 33.]) was used to transform the molecular structure files, and CSD Python API (Moghadam et al., 2020[Moghadam, P. Z., Li, A., Liu, X. W., Bueno-Perez, R., Wang, S. D., Wiggin, S. B., Wood, P. A. & Fairen-Jimenez, D. (2020). Chem. Sci. 11, 8373-8387.]) was used to obtain and manipulate the crystal structures. The model training was performed on a Windows 10 OS workstation with an Intel Xeon E5-2620 v3 (12 cores, 24 threads) CPU, 128 GB RAM and a GPU (NVIDIA Tesla K40, 12GB).

3. Results and discussion

3.1. Dataset overview

A histogram of the metal complexes is presented in Fig. 3[link] to determine the composition of the molecules included in the dataset. This classification is based on seven categories: (i) complexes consisting of a single lanthanide ion, (ii) homonuclear lanthanide ion binuclear complexes, (iii) homonuclear transition metal binuclear complexes, (iv) 3d–4f complexes, (v) 4d–4f or 5d–4f complexes, (vi) complexes consisting of a single transition metal, and (vii) actinide ions (uranium). Histograms illustrating the number of atoms and nuclei in the SMMs and non-SMMs are provided in the supporting information (Figs. S1 and S2).

[Figure 3]
Figure 3
Histogram of SMM and non-SMM molecules included in the dataset. Blue indicates SMM molecules, and orange indicates non-SMM molecules. (1) Complexes consisting of a single lanthanoid ion. (2) Homonuclear lanthanide ion dinuclear complexes. (3) Homonuclear transition metal dinuclear complexes. (4) 3d–4f complexes. (5) 4d–4f complexes or 5d–4f complexes. (6) Complexes consisting of a single transition metal. (7) Complexes containing actinide ions (uranium).

Figs. 3[link], S1 and 2 indicate that the differences were not pronounced depending on the number of atoms or nuclei. Histograms depicting complexes containing lanthanide ions show that those containing Dy ions display a significant proportion of SMM properties (Fig. S2a), consistent with several previous results (Zhang et al., 2013[Zhang, P., Guo, Y.-N. & Tang, J. (2013). Coord. Chem. Rev. 257, 1728-1763.]; Dey et al., 2018[Dey, A., Kalita, P. & Chandrasekhar, V. (2018). ACS Omega, 3, 9462-9475.]). Complexes containing Dy, Gd and Tb ions constituted a significant proportion of the data. However, for the other lanthanide ions, the quantity of data was minimal, and no distinctions in the SMM properties due to elemental variations were discernible. For the 3d–4f complexes, Dy–Zn complexes are the most prevalent molecules with SMM properties, and Dy–Cu, Dy–Ni and Tb–Zn complexes are also included. Alternatively, Dy–d complexes are still prevalent in molecules lacking SMM properties. However, note that Tb–Cu complexes were the most common (Figs. S3 and S4). It is challenging to use traditional machine learning to forecast SMM properties because of the difficulty in identifying differences by element or number of nuclei and the presence of Dy complexes that do not exhibit SMM properties.

3.2. Learning results

The learning curve is illustrated in Fig. S5, indicating positive progress as the loss decreases with learning. The similarity between the training and generalization errors suggests that the deep-learning model is robust to unknown data. The final accuracy and AUC were approximately 70% (Table 1[link]). The confusion matrix illustrated in Fig. 4[link] exhibits substantial values for the diagonal components. Distinguishing SMMs from molecular structures alone is challenging, even for the human eye. Notably, the deep-learning model demonstrated 70% accuracy, and the prediction in this study was based solely on voxel information generated from the type of element and its positional relationship.

Table 1
Correct answer rate and AUC on training, validation and test data

  ACC AUC
Train 0.7250 0.8185
Valid 0.6875 0.7758
Test 0.7000 0.7269
[Figure 4]
Figure 4
Confusion matrix in test data.

It is expected that enhancing the descriptors and refining the learning model will result in improved accuracy. In particular, utilizing spin and orbital angular momentum is crucial in magnetism. Previous studies have employed charge distributions and spin densities to obtain voxel information (Kuzminykh et al., 2018[Kuzminykh, D., Polykovskiy, D., Kadurin, A., Zhebrak, A., Baskov, I., Nikolenko, S., Shayakhmetov, R. & Zhavoronkov, A. (2018). Mol. Pharm. 15, 4378-4385.]; Ghosh et al., 2019[Ghosh, K., Stuke, A., Todorović, M., Jørgensen, P. B., Schmidt, M. N., Vehtari, A. & Rinke, P. (2019). Adv. Sci. 6, 1970053.]; Casey et al., 2020[Casey, A. D., Son, S. F., Bilionis, I. & Barnes, B. C. (2020). J. Chem. Inf. Model. 60, 4457-4473.]). Quantum chemical calculations can determine these values. However, such calculations require significant computational time and resources when dealing with chemical species containing transition and lanthanide metals. These calculations are unfeasible for studies involving hundreds or tens of thousands of molecules, like in this study. In addition, it would be desirable to compare the performance with the results obtained using graphs, molecular descriptors and voxels. However, numerous methods are inadequate for molecules containing metal ions, necessitating the renovation of programs. Unfortunately, this action was not performed in the current investigation because of time restrictions. Further studies are required to address this issue.

3.3. Visualization of learning results

Grad-CAM (Selvaraju et al., 2017[Selvaraju, R. R., Cogswell, M., Das, A., Vedantam, R., Parikh, D. & Batra, D. (2017). Proceedings of the 2017 IEEE Conference on Computer Vision and Pattern Recognition, 21- 26 July 2017, Honolulu, Hawaii, USA, pp. 618-626. Washington DC: IEEE Press.]) was used to visualize the area of focus of the model for the inputs. The Grad-CAM diagram indicates that the model is concentrated near the central metal (Fig. 5[link]). Notably, the model recognizes that the central metal initiates magnetism in metal complexes. Fig. S6 demonstrates the successful and unsuccessful predictions, and the model does not focus on significantly misguided areas, even in failed predictions. Furthermore, this model is not expected to focus on particular ligands or functional groups to perform optimally in unfamiliar molecular structures.

[Figure 5]
Figure 5
Visualization of CNN prediction results using Grad-CAM. The red blocks in the image are the parts that make a significant contribution when CNN makes predictions. It is obtained as voxels 64 × 64 × 64 in size, but was divided into 64 2D images of 64 × 64 for plotting.

In contrast, we expected to discover new structural features that would go beyond the thinking of chemists; however, for this study, we could only obtain the same insights. The opaque nature of deep-learning predictions impedes thorough analysis, leading to tentative conclusions. However, commonly employed chemoinformatics techniques [graph-neural networks (GNNs) and recurrent neural networks (RNNs)] frequently lack visualization methods compared with CNNs. This attribute of CNNs is advantageous for property prediction, rendering it a potent technique in fields lacking a comprehensive understanding of the reaction mechanism, for example in protein–ligand docking and catalysis studies.

3.4. Proof of SMM search with the CSD

Magnetic measurements were conducted in the solid state; therefore, the SMM molecular design must account for the molecular structure within the crystal structure. Predicting the crystal structure of the metal complexes is necessary to design SMM molecules from scratch. However, a method to achieve this goal is yet to be established. Fortunately, the CSD contains one million crystal structures, of which only a small proportion have undergone AC susceptibility measurements. Predicting the SMM behaviour of the crystal structures registered in the CSD was validated as a model case for designing SMM molecules using deep learning.

Approximately 20 000 crystal structures of metal complexes containing Schiff bases were obtained using the program ConQuest. Using the CSD Python API, the molecular structures of the metal complexes were extracted from the crystal structures by removing crystalline water and other components. The SMM behaviours of the molecular structures predicted by the learned model are shown in Fig. 6[link]. The focus was on the top 10 molecules predicted to be the most SMM-like and the bottom 10 predicted to be the least SMM-like. The original publications were searched using CCDC numbers, and the pivotal metal and magnetic measurements are summarized in Tables 2[link] and 3[link]. Numerous molecules predicted to be SMMs were confirmed to exhibit SMM behaviour. Conversely, many molecules predicted to be non-SMMs have not been measured for AC susceptibility.

Table 2
The top 10 molecules predicted most likely to be SMM

`No' if reported not, and `–' if AC susceptibility measurements were not performed.

  CCDC Score SMM Metal Reference
1 QURMUN 0.0076 Yes Dy18Zn12 Stavgianoudaki et al. (2016[Stavgianoudaki, N., Siczek, M., Lis, T., Inglis, R. & Milios, C. J. (2016). Chem. Commun. 52, 343-345.])
2 VUVKEE 0.0189 No Dy10 Das et al. (2015[Das, S., Dey, A., Kundu, S., Biswas, S., Narayanan, R. S., Titos-Padilla, S., Lorusso, G., Evangelisti, M., Colacio, E. & Chandrasekhar, V. (2015). Chem. A Eur. J. 21, 16955-16967.])
3 REYYEB 0.0228 Yes Dy5Mn4 Alexandropoulos et al. (2013[Alexandropoulos, D. I., Nguyen, T. N., Cunha-Silva, L., Zafiropoulos, T. F., Escuer, A., Christou, G. & Stamatatos, T. C. (2013). Inorg. Chem. 52, 1179-1181.])
4 RECPIB 0.0289 Yes Dy6 Lin et al. (2017[Lin, S.-Y., Wernsdorfer, W., Ungur, L., Powell, A. K., Guo, Y.-N., Tang, J., Zhao, L., Chibotaru, L. F. & Zhang, H.-J. (2012). Angew. Chem. Int. Ed. 51, 12767-12771.])
5 ZEZYIO 0.0327 Yes Dy6 Lin et al. (2012[Lin, S.-Y., Wu, J. & Xu, Z. (2017). RSC Adv. 7, 47520-47526.])
6 LOSVOG 0.0331 Yes Dy9 Zou et al. (2015[Zou, H.-H., Sheng, L.-B., Chen, Z.-L. & Liang, F.-P. (2015). Polyhedron, 88, 110-115.])
7 OPEFOH 0.0441 No Dy6 Schlittenhardt et al. (2021[Schlittenhardt, S., Moreno-Pineda, E. & Ruben, M. (2021). Dalton Trans. 50, 4735-4742.])
8 EZEZUG 0.0453 No Dy12 Li et al. (2016[Li, Y.-M., Kuang, W.-W., Zhu, L.-L., Xu, Y. & Yang, P.-P. (2016). Eur. J. Inorg. Chem. 2016, 4996-5003.])
9 YALRAF 0.0472 UCs Cametti et al. (2005[Cametti, M., Nissinen, M., Dalla Cort, A., Mandolini, L. & Rissanen, K. (2005). J. Am. Chem. Soc. 127, 3831-3837.])
10 OPEGAU 0.0475 No Dy6 Schlittenhardt et al. (2021[Schlittenhardt, S., Moreno-Pineda, E. & Ruben, M. (2021). Dalton Trans. 50, 4735-4742.])
†Material exhibits magnetic relaxation behaviour under a DC magnetic field.

Table 3
The bottom 10 molecules predicted least likely to be SMM

`No' if reported not, and `–' if AC susceptibility measurements were not performed.

  CCDC Score SMM Metal Reference
1 PIVGAD 0.9990 Nd2Cu2 Gheorghe et al. (2007[Gheorghe, R., Andruh, M., Costes, J.-P., Donnadieu, B., Schmidtmann, M. & Müller, A. (2007). Inorg. Chim. Acta, 360, 4044-4050. ])
2 TOKRAO 0.9990 No Ni15 Muche et al. (2014[Muche, S., Levacheva, I., Samsonova, O., Pham, L., Christou, G., Bakowsky, U. & Hołyńska, M. (2014). Inorg. Chem. 53, 7642-7649. ])
3 VICQUV 0.9991 Nd2 Zou et al. (2013[Zou, X., Li, M., Yan, P., Zhang, J., Hou, G. & Li, G. (2013). Dalton Trans. 42, 9482-9489.])
4 VAKLEY 0.9994 Nd Ling et al. (1989[Ling, H., Chunhui, H. & Guangxian, X. (1989). Jiegou Huaxue, 8, 1.])
5 YUTCAS 0.9994 Nd2Zn2 et al. (2010[Lü, X. Q., Feng, W. X., Hui, Y. N., Wei, T., Song, J. R., Zhao, S. S., Wong, W. Y., Wong, W. K. & Jones, R. A. (2010). Eur. J. Inorg. Chem. 2010, 2714-2722.])
6 YEBPOL 0.9994 PrCu2 Gheorghe et al. (2006[Gheorghe, R., Cucos, P., Andruh, M., Costes, J.-P., Donnadieu, B. & Shova, S. (2006). Chem. A Eur. J. 12, 187-203.])
7 XUXNEK 0.9995 Pr2Cu2 Pointillart et al. (2010[Pointillart, F. & Bernot, K. (2010). Eur. J. Inorg. Chem. 2010, 952-964.])
8 SIGZIU 0.9995 Nd3Zn6 Song et al. (2018[Song, X. Q., Wang, C. Y., Meng, H. H., Shamshoom, A. A. A. & Liu, W. S. (2018). Inorg. Chem. 57, 10873-10880.])
9 SIGZEQ 0.9996 Nd8Zn12 Song et al. (2018[Song, X. Q., Wang, C. Y., Meng, H. H., Shamshoom, A. A. A. & Liu, W. S. (2018). Inorg. Chem. 57, 10873-10880.])
10 TIYGUF 0.9998 Nd6Cd18 Yang et al. (2013[Yang, X., Schipper, D., Jones, R. A., Lytwak, L. A., Holliday, B. J. & Huang, S. (2013). J. Am. Chem. Soc. 135, 8468-8471.])
[Figure 6]
Figure 6
Histogram of predicted SMM properties for approximately 20 000 molecules obtained from the CSD. The closer the frequency is to 0, the more SMM-like it is predicted to be.

Many of the predicted SMM molecules had structures comprising multinuclear Dy complexes. According to equation (1)[link], a larger total number of spins in the complex leads to a more significant value of Ueff. This approach is similar to earlier approaches that enhance the spin multiplicity S using multinuclear complexes of d-metals. Previous studies have shown that symmetry reduction occurs with increased nuclei, and the highest Ueff value was 40 K for ZEZYIO (Lin et al., 2012[Lin, S.-Y., Wernsdorfer, W., Ungur, L., Powell, A. K., Guo, Y.-N., Tang, J., Zhao, L., Chibotaru, L. F. & Zhang, H.-J. (2012). Angew. Chem. Int. Ed. 51, 12767-12771.]). This study conducts learning based on the presence or absence of SMM behaviour, with no consideration given to the height of Ueff for simplicity. However, to identify the more promising SMMs, a dataset that considers Ueff should be able to select the best SMMs.

It is important to emphasize that the molecular structure data for the 20 000 molecules used in this study did not include the data used in training. Moreover, while salen-type ligands were prominent in the training phase, a more comprehensive range of ligands (Schiff bases) was included in the 20 000 data points. Thus, the deep-learning model solely predicted unknown molecules without knowing the answers, ultimately selecting the SMM molecules from the CSD. For instance, novel SMM molecules may be uncovered by re-synthesizing molecules that have been predicted to be SMMs but have not been subjected to AC susceptibility and magnetic measurements. The deep-learning-based SMM research model established in this study serves as a practical data-driven chemistry framework.

As referenced in previous research (Sumita et al., 2018[Sumita, M., Yang, X., Ishihara, S., Tamura, R. & Tsuda, K. (2018). ACS Cent. Sci. 4, 1126-1133.], 2022[Sumita, M., Terayama, K., Suzuki, N., Ishihara, S., Tamura, R., Chahal, M. K., Payne, D. T., Yoshizoe, K. & Tsuda, K. (2022). Sci. Adv. 8, eabj3906.]), for AI-assisted molecular design, it is recommended to use data-independent approaches that involve both (1) AI-generated molecular structures and (2) prediction of properties via quantum chemical calculations. AI generates extensive candidate molecules, ranging from tens to hundreds of millions, and property predictions are performed via large-scale supercomputing. The latter is facilitated using supercomputers to predict computational properties. Electronic structure calculations of organic molecules in solution are highly accurate and have successfully controlled UV absorption and created fluorescent molecules (Sumita et al., 2018[Sumita, M., Yang, X., Ishihara, S., Tamura, R. & Tsuda, K. (2018). ACS Cent. Sci. 4, 1126-1133.], 2022[Sumita, M., Terayama, K., Suzuki, N., Ishihara, S., Tamura, R., Chahal, M. K., Payne, D. T., Yoshizoe, K. & Tsuda, K. (2022). Sci. Adv. 8, eabj3906.]). However, solid-state studies on the magnetism of metal complexes encounter two challenges. The first is the obstacle in predicting the crystal structures of metal complexes (Desiraju, 2013[Desiraju, G. R. (2013). J. Am. Chem. Soc. 135, 9952-9967.]; McDonagh et al., 2019[McDonagh, D., Skylaris, C.-K. & Day, G. M. (2019). J. Chem. Theory Comput. 15, 2743-2758.]). The second is the challenge of simulating magnetism. Novel breakthroughs must be made to resolve these issues. As a compromise, this study designs SMM molecules using molecular structures from a crystal database and data-driven prediction by deep learning. However, the data are limited, and the accuracy of the results depends on the quantity of available data. Moreover, because the results were data-dependent, they were not significantly superior to those obtained by conventional molecular design by conventional chemists. Furthermore, it is not feasible to acquire innovative SMMs solely by predicting molecules with well defined crystal structures. Further research is required in the future.

4. Conclusions

This study conducted deep learning using experimental data from SMM studies and molecular structures derived from crystal structures. The deep-learning model successfully acquired knowledge of the structural features of the SMM and accurately predicted the SMM behaviour using only 3D structure and magnetic data. We have demonstrated that deep-learning models can identify SMM molecules within the CSD. This study presents a data-driven approach to experimentation instead of relying on subjective evaluations based on intuition and experience. With further developments and the integration of generative AI and first-principles calculations, unprecedented SMMs can be achieved using deep-learning methods.

Supporting information


References

First citationAbadi, M., Barham, P., Chen, J., Chen, Z., Davis, A., Dean, J., Devin, M., Ghemawat, S., Irving, G. & Isard, M. (2016). Proceedings of the 12th USENIX Symposium on Operating Systems Design and Implementation (OSDI16), 2–4 November 2016, Savannah, GA, USA, pp. 265–283. The USENIX Association.  Google Scholar
First citationAgarap, A. F. (2019). Deep learning using rectified linear units (ReLU), arXiv: 1803.08375.  Google Scholar
First citationAlexandropoulos, D. I., Nguyen, T. N., Cunha-Silva, L., Zafiropoulos, T. F., Escuer, A., Christou, G. & Stamatatos, T. C. (2013). Inorg. Chem. 52, 1179–1181.  Web of Science CSD CrossRef CAS PubMed Google Scholar
First citationAmidi, A., Amidi, S., Vlachakis, D., Megalooikonomou, V., Paragios, N. & Zacharaki, E. I. (2018). PeerJ, 6, e4750.  Web of Science CrossRef PubMed Google Scholar
First citationAquilante, F., De Vico, L., Ferré, N., Ghigo, G., Malmqvist, P., Neogrády, P., Pedersen, T. B., Pitoňák, M., Reiher, M., Roos, B. O., Serrano–Andrés, L., Urban, M., Veryazov, V. & Lindh, R. (2010). J. Comput. Chem. 31, 224–247.  Web of Science CrossRef PubMed CAS Google Scholar
First citationAravena, D. (2018). J. Phys. Chem. Lett. 9, 5327–5333.  Web of Science CrossRef CAS PubMed Google Scholar
First citationBishop, C. M. (1995). Neural Networks for Pattern Recognition. Oxford University Press.  Google Scholar
First citationBjorck, N., Gomes, C. P., Selman, B. & Weinberger, K. Q. (2018). Advances in Neural Information Processing Systems 31 (NeurIPS 2018), 3–8 December 2018, Montreal, Canada, pp. 31.  Google Scholar
First citationCametti, M., Nissinen, M., Dalla Cort, A., Mandolini, L. & Rissanen, K. (2005). J. Am. Chem. Soc. 127, 3831–3837.  Web of Science CSD CrossRef PubMed CAS Google Scholar
First citationCasey, A. D., Son, S. F., Bilionis, I. & Barnes, B. C. (2020). J. Chem. Inf. Model. 60, 4457–4473.  Web of Science CrossRef CAS PubMed Google Scholar
First citationCeriotti, M., Clementi, C. & Anatole von Lilienfeld, O. (2021). Chem. Rev. 121, 9719–9721.  Web of Science CrossRef CAS PubMed Google Scholar
First citationCybenko, G. (1989). Math. Contr. Signal. Syst. 2, 303–314.  CrossRef Google Scholar
First citationDas, S., Dey, A., Kundu, S., Biswas, S., Narayanan, R. S., Titos–Padilla, S., Lorusso, G., Evangelisti, M., Colacio, E. & Chandrasekhar, V. (2015). Chem. A Eur. J. 21, 16955–16967.  Web of Science CSD CrossRef CAS Google Scholar
First citationDesiraju, G. R. (2013). J. Am. Chem. Soc. 135, 9952–9967.  Web of Science CrossRef CAS PubMed Google Scholar
First citationDey, A., Kalita, P. & Chandrasekhar, V. (2018). ACS Omega, 3, 9462–9475.  Web of Science CrossRef CAS PubMed Google Scholar
First citationDuan, Y., Rosaleny, L. E., Coutinho, J. T., Giménez-Santamarina, S., Scheie, A., Baldoví, J. J., Cardona-Serra, S. & Gaita-Ariño, A. (2022). Nat. Commun. 13, 7626.  Web of Science CrossRef PubMed Google Scholar
First citationEliseeva, S. V. & Bünzli, J. G. (2011). New J. Chem. 35, 1165–1176.  Web of Science CrossRef CAS Google Scholar
First citationElton, D. C., Boukouvalas, Z., Fuge, M. D. & Chung, P. W. (2019). Mol. Syst. Des. Eng. 4, 828–849.  Web of Science CrossRef CAS Google Scholar
First citationEngel, T. (2006). J. Chem. Inf. Model. 46, 2267–2277.  Web of Science CrossRef PubMed CAS Google Scholar
First citationGal, Y. & Ghahramani, Z. (2015). arXiv: 1506.02158.  Google Scholar
First citationGatteschi, D., Sessoli, R. & Villain, J. (2006). Molecular Nanomagnets, Oxford University Press.  Google Scholar
First citationGheorghe, R., Andruh, M., Costes, J.-P., Donnadieu, B., Schmidtmann, M. & Müller, A. (2007). Inorg. Chim. Acta, 360, 4044–4050.   Web of Science CSD CrossRef CAS Google Scholar
First citationGheorghe, R., Cucos, P., Andruh, M., Costes, J.-P., Donnadieu, B. & Shova, S. (2006). Chem. A Eur. J. 12, 187–203.  Web of Science CSD CrossRef CAS Google Scholar
First citationGhosh, K., Stuke, A., Todorović, M., Jørgensen, P. B., Schmidt, M. N., Vehtari, A. & Rinke, P. (2019). Adv. Sci. 6, 1970053.  CrossRef Google Scholar
First citationGroom, C. R., Bruno, I. J., Lightfoot, M. P. & Ward, S. C. (2016). Acta Cryst. B72, 171–179.  Web of Science CrossRef IUCr Journals Google Scholar
First citationHarris, C. R., Millman, K. J., van der Walt, S. J., Gommers, R., Virtanen, P., Cournapeau, D., Wieser, E., Taylor, J., Berg, S., Smith, N. J., Kern, R., Picus, M., Hoyer, S., van Kerkwijk, M. H., Brett, M., Haldane, A., del Río, J. F., Wiebe, M., Peterson, P., Gérard-Marchant, P., Sheppard, K., Reddy, T., Weckesser, W., Abbasi, H., Gohlke, C. & Oliphant, T. E. (2020). Nature, 585, 357–362.  Web of Science CrossRef CAS PubMed Google Scholar
First citationHe, K., Zhang, X., Ren, S. & Sun, J. (2015). Proceedings of the IEEE conference on computer vision and Pattern Recognition (CVPR2015), 7–12 June 2024, Boston, MA, USA, pp. 70–778. IEEE Press.  Google Scholar
First citationJanet, J. P., Liu, F., Nandy, A., Duan, C., Yang, T., Lin, S. & Kulik, H. J. (2019). Inorg. Chem. 58, 10592–10606.  Web of Science CrossRef CAS PubMed Google Scholar
First citationJogin, M., Madhulika, M., Divya, G., Meghana, R. & Apoorva, S. (2018). Proceedings of the 3rd IEEE International Conference on Recent Trends in Electronics, Information & Communication Technology (RTEICT), 18–19 May 2018, Bengaluru, India, pp. 2319–2323. IEEE Press.  Google Scholar
First citationKuzminykh, D., Polykovskiy, D., Kadurin, A., Zhebrak, A., Baskov, I., Nikolenko, S., Shayakhmetov, R. & Zhavoronkov, A. (2018). Mol. Pharm. 15, 4378–4385.  Web of Science CrossRef CAS PubMed Google Scholar
First citationLeuenberger, M. N. & Loss, D. (2001). Nature, 410, 789–793.  Web of Science CrossRef PubMed CAS Google Scholar
First citationLi, C., Wang, J., Niu, Z., Yao, J. & Zeng, X. (2021). Brief. Bioinform. 22, bbab078.  Web of Science CrossRef PubMed Google Scholar
First citationLi, Y.-M., Kuang, W.-W., Zhu, L.-L., Xu, Y. & Yang, P.-P. (2016). Eur. J. Inorg. Chem. 2016, 4996–5003.  Web of Science CSD CrossRef CAS Google Scholar
First citationLin, S.-Y., Wernsdorfer, W., Ungur, L., Powell, A. K., Guo, Y.-N., Tang, J., Zhao, L., Chibotaru, L. F. & Zhang, H.-J. (2012). Angew. Chem. Int. Ed. 51, 12767–12771.  Web of Science CSD CrossRef CAS Google Scholar
First citationLin, S.-Y., Wu, J. & Xu, Z. (2017). RSC Adv. 7, 47520–47526.  Web of Science CSD CrossRef CAS Google Scholar
First citationLing, H., Chunhui, H. & Guangxian, X. (1989). Jiegou Huaxue, 8, 1.  Google Scholar
First citationLiu, K., Shi, W. & Cheng, P. (2015). Coord. Chem. Rev. 289–290, 74–122.  Web of Science CrossRef CAS Google Scholar
First citationLoshchilov, I. & Hutter, F. (2016). arXiv: 1608.03983.  Google Scholar
First citationLü, X. Q., Feng, W. X., Hui, Y. N., Wei, T., Song, J. R., Zhao, S. S., Wong, W. Y., Wong, W. K. & Jones, R. A. (2010). Eur. J. Inorg. Chem. 2010, 2714–2722.  Google Scholar
First citationMacrae, C. F., Sovago, I., Cottrell, S. J., Galek, P. T. A., McCabe, P., Pidcock, E., Platings, M., Shields, G. P., Stevens, J. S., Towler, M. & Wood, P. A. (2020). J. Appl. Cryst. 53, 226–235.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationMcDonagh, D., Skylaris, C.-K. & Day, G. M. (2019). J. Chem. Theory Comput. 15, 2743–2758.  Web of Science CrossRef CAS PubMed Google Scholar
First citationMitchell, J. B. (2014). WIREs Comput. Mol. Sci. 4, 468–481.  Web of Science CrossRef CAS Google Scholar
First citationMoghadam, P. Z., Li, A., Liu, X. W., Bueno-Perez, R., Wang, S. D., Wiggin, S. B., Wood, P. A. & Fairen-Jimenez, D. (2020). Chem. Sci. 11, 8373–8387.  Web of Science CrossRef CAS PubMed Google Scholar
First citationMohammed, R., Rawashdeh, J. & Abdullah, M. (2020). Proceedings of the 11th International Conference on Information and Communications Systems (ICICS2020), 7–9 April 2020, Irbid, Jordan, pp. 243–248. IEEE Press.  Google Scholar
First citationMuche, S., Levacheva, I., Samsonova, O., Pham, L., Christou, G., Bakowsky, U. & Hołyńska, M. (2014). Inorg. Chem. 53, 7642–7649.   Web of Science CSD CrossRef CAS PubMed Google Scholar
First citationNandy, A., Duan, C., Taylor, M. G., Liu, F., Steeves, A. H. & Kulik, H. J. (2021). Chem. Rev. 121, 9927–10000.  Web of Science CrossRef CAS PubMed Google Scholar
First citationO'Boyle, N. M., Banck, M., James, C. A., Morley, C., Vandermeersch, T. & Hutchison, G. R. (2011). J. Cheminform, 3, 33.  Web of Science PubMed Google Scholar
First citationPark, S. & Seok, C. (2022). J. Chem. Inf. Model. 62, 3157–3168.  Web of Science CrossRef CAS PubMed Google Scholar
First citationParmar, V. S., Mills, D. P. & Winpenny, R. E. P. (2021). Chem. A Eur. J. 27, 7625–7645.  Web of Science CrossRef CAS Google Scholar
First citationPointillart, F. & Bernot, K. (2010). Eur. J. Inorg. Chem. 2010, 952–964.  Web of Science CSD CrossRef Google Scholar
First citationRandall McClain, K. R., Gould, C. A., Chakarawet, K., Teat, S. J., Groshens, T. J., Long, J. R. & Harvey, B. G. (2018). Chem. Sci. 9, 8492–8503.  Web of Science CSD CrossRef CAS PubMed Google Scholar
First citationReddi, S. J., Kale, S. & Kumar, S. (2019). arXiv: 1904.09237.  Google Scholar
First citationRoos, B. O.-Å. & Malmqvist, P. (2004). Phys. Chem. Chem. Phys. 6, 2919–2927.  Web of Science CrossRef CAS Google Scholar
First citationRyan, K., Lengyel, J. & Shatruk, M. (2018). J. Am. Chem. Soc. 140, 10158–10168.  Web of Science CrossRef PubMed Google Scholar
First citationSchlittenhardt, S., Moreno-Pineda, E. & Ruben, M. (2021). Dalton Trans. 50, 4735–4742.  Web of Science CSD CrossRef CAS PubMed Google Scholar
First citationSelvaraju, R. R., Cogswell, M., Das, A., Vedantam, R., Parikh, D. & Batra, D. (2017). Proceedings of the 2017 IEEE Conference on Computer Vision and Pattern Recognition, 21– 26 July 2017, Honolulu, Hawaii, USA, pp. 618–626. Washington DC: IEEE Press.  Google Scholar
First citationSessoli, R., Gatteschi, D., Caneschi, A. & Novak, M. A. (1993). Nature, 365, 141–143.  CrossRef CAS Web of Science Google Scholar
First citationShorten, C. & Khoshgoftaar, T. M. (2019). J. Big Data, 6, 60.  Web of Science CrossRef Google Scholar
First citationSilver, D., Huang, A., Maddison, C. J., Guez, A., Sifre, L., van den Driessche, G., Schrittwieser, J., Antonoglou, I., Panneershelvam, V., Lanctot, M., Dieleman, S., Grewe, D., Nham, J., Kalchbrenner, N., Sutskever, I., Lillicrap, T., Leach, M., Kavukcuoglu, K., Graepel, T. & Hassabis, D. (2016). Nature, 529, 484–489.  Web of Science CrossRef CAS PubMed Google Scholar
First citationSong, X. Q., Wang, C. Y., Meng, H. H., Shamshoom, A. A. A. & Liu, W. S. (2018). Inorg. Chem. 57, 10873–10880.  Web of Science CSD CrossRef CAS PubMed Google Scholar
First citationStavgianoudaki, N., Siczek, M., Lis, T., Inglis, R. & Milios, C. J. (2016). Chem. Commun. 52, 343–345.  Web of Science CSD CrossRef CAS Google Scholar
First citationSumita, M., Terayama, K., Suzuki, N., Ishihara, S., Tamura, R., Chahal, M. K., Payne, D. T., Yoshizoe, K. & Tsuda, K. (2022). Sci. Adv. 8, eabj3906.  Web of Science CrossRef PubMed Google Scholar
First citationSumita, M., Yang, X., Ishihara, S., Tamura, R. & Tsuda, K. (2018). ACS Cent. Sci. 4, 1126–1133.  Web of Science CrossRef CAS PubMed Google Scholar
First citationTang, J. & Zhang, P. (2015). Lanthanide Single Molecule Magnets, 2 Berlin: Springer.  Google Scholar
First citationUngur, L. & Chibotaru, L. F. (2011). Phys. Chem. Chem. Phys. 13, 20086–20090.  Web of Science CrossRef CAS PubMed Google Scholar
First citationVirtanen, P., Gommers, R., Oliphant, T. E., Haberland, M., Reddy, T., Cournapeau, D., Burovski, E., Peterson, P., Weckesser, W., Bright, J., van der Walt, S. J., Brett, M., Wilson, J., Millman, K. J., Mayorov, N., Nelson, A. R. J., Jones, E., Kern, R., Larson, E., Carey, C. J., Polat, İ., Feng, Y., Moore, E. W., VanderPlas, J., Laxalde, D., Perktold, J., Cimrman, R., Henriksen, I., Quintero, E. A., Harris, C. R., Archibald, A. M., Ribeiro, A. H., Pedregosa, F., van Mulbregt, P., Vijaykumar, A., Bardelli, A. P., Rothberg, A., Hilboll, A., Kloeckner, A., Scopatz, A., Lee, A., Rokem, A., Woods, C. N., Fulton, C., Masson, C., Häggström, C., Fitzgerald, C., Nicholson, D. A., Hagen, D. R., Pasechnik, D. V., Olivetti, E., Martin, E., Wieser, E., Silva, F., Lenders, F., Wilhelm, F., Young, G., Price, G. A., Ingold, G., Allen, G. E., Lee, G. R., Audren, H., Probst, I., Dietrich, J. P., Silterra, J., Webber, J. T., Slavič, J., Nothman, J., Buchner, J., Kulick, J., Schönberger, J. L., de Miranda Cardoso, J. V., Reimer, J., Harrington, J., Rodríguez, J. L. C., Nunez-Iglesias, J., Kuczynski, J., Tritz, K., Thoma, M., Newville, M., Kümmerer, M., Bolingbroke, M., Tartre, M., Pak, M., Smith, N. J., Nowaczyk, N., Shebanov, N., Pavlyk, O., Brodtkorb, P. A., Lee, P., McGibbon, R. T., Feldbauer, R., Lewis, S., Tygier, S., Sievert, S., Vigna, S., Peterson, S., More, S., Pudlik, T., Oshima, T., Pingel, T. J., Robitaille, T. P., Spura, T., Jones, T. R., Cera, T., Leslie, T., Zito, T., Krauss, T., Upadhyay, U., Halchenko, Y. O. & Vázquez-Baeza, Y. (2020). Nat. Methods, 17, 261–272.  Web of Science CrossRef CAS PubMed Google Scholar
First citationWinpenny, R. & Aromí, G. (2006). Single-Molecule Magnets and Related Phenomena, p. 122. Berlin: Springer.  Google Scholar
First citationWoodruff, D. N., Winpenny, R. E. & Layfield, R. A. (2013). Chem. Rev. 113, 5110–5148.  Web of Science CrossRef CAS PubMed Google Scholar
First citationYang, X., Schipper, D., Jones, R. A., Lytwak, L. A., Holliday, B. J. & Huang, S. (2013). J. Am. Chem. Soc. 135, 8468–8471.  Web of Science CSD CrossRef CAS PubMed Google Scholar
First citationYin, B. & Li, C. C. (2020). Phys. Chem. Chem. Phys. 22, 9923–9933.  Web of Science CrossRef CAS PubMed Google Scholar
First citationZhang, P., Guo, Y.-N. & Tang, J. (2013). Coord. Chem. Rev. 257, 1728–1763.  Web of Science CrossRef CAS Google Scholar
First citationZou, H.-H., Sheng, L.-B., Chen, Z.-L. & Liang, F.-P. (2015). Polyhedron, 88, 110–115.  Web of Science CSD CrossRef CAS Google Scholar
First citationZou, X., Li, M., Yan, P., Zhang, J., Hou, G. & Li, G. (2013). Dalton Trans. 42, 9482–9489.  Web of Science CSD CrossRef CAS PubMed Google Scholar

This is an open-access article distributed under the terms of the Creative Commons Attribution (CC-BY) Licence, which permits unrestricted use, distribution, and reproduction in any medium, provided the original authors and source are cited.

IUCrJ
Volume 11| Part 2| March 2024| Pages 182-189
ISSN: 2052-2525