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

Journal logoSTRUCTURAL SCIENCE
CRYSTAL ENGINEERING
MATERIALS
ISSN: 2052-5206

Contrasting conformational behaviors of molecules XXXI and XXXII in the seventh blind test of crystal structure prediction

crossmark logo

aDepartment of Chemistry, University of California, Riverside, CA 92521, USA
*Correspondence e-mail: gregory.beran@ucr.edu

Edited by A. Nangia, CSIR–National Chemical Laboratory, India (Received 5 February 2024; accepted 29 May 2024; online 14 October 2024)

This article is part of a collection of articles covering the seventh crystal structure prediction blind test.

Accurate modeling of conformational energies is key to the crystal structure prediction of conformational polymorphs. Focusing on molecules XXXI and XXXII from the seventh blind test of crystal structure prediction, this study employs various electronic structure methods up to the level of domain-local pair natural orbital coupled cluster singles and doubles with perturbative triples [DLPNO-CCSD(T1)] to benchmark the conformational energies and to assess their impact on the crystal energy landscapes. Molecule XXXI proves to be a relatively straightforward case, with the conformational energies from generalized gradient approximation (GGA) functional B86bPBE-XDM changing only modestly when using more advanced density functionals such as PBE0-D4, ωB97M-V, and revDSD-PBEP86-D4, dispersion-corrected second-order Møller–Plesset perturbation theory (SCS-MP2D), or DLPNO-CCSD(T1). In contrast, the conformational energies of molecule XXXII prove difficult to determine reliably, and variations in the computed conformational energies appreciably impact the crystal energy landscape. Even high-level methods such as revDSD-PBEP86-D4 and SCS-MP2D exhibit significant disagreements with the DLPNO-CCSD(T1) benchmarks for molecule XXXII, highlighting the difficulty of predicting conformational energies for complex, drug-like molecules. The best-converged predicted crystal energy landscape obtained here for molecule XXXII disagrees significantly with what has been inferred about the solid-form landscape experimentally. The identified limitations of the calculations are probably insufficient to account for the discrepancies between theory and experiment on molecule XXXII, and further investigation of the experimental solid-form landscape would be valuable. Finally, assessment of several semi-empirical methods finds r2SCAN-3c to be the most promising, with conformational energy accuracy intermediate between the GGA and hybrid functionals and a low computational cost.

1. Introduction

Over the past few decades, organic molecular crystal structure prediction (CSP) has transformed from a seemingly intractable problem to one where successful predictions are common. CSP is increasingly being used to understand pharmaceutical solid form landscapes, to help solve challenging crystal structures, and to help design new organic materials (Beran, 2023[Beran, G. J. O. (2023). Chem. Sci. 14, 13290-13312.]). The seven blind tests of crystal structure prediction between 1999 and 2022 have played a central role in highlighting the current capabilities of the models being used and the future improvements needed (Lommerse et al., 2000[Lommerse, J. P. M., Motherwell, W. D. S., Ammon, H. L., Dunitz, J. D., Gavezzotti, A., Hofmann, D. W. M., Leusen, F. J. J., Mooij, W. T. M., Price, S. L., Schweizer, B., Schmidt, M. U., van Eijck, B. P., Verwer, P. & Williams, D. E. (2000). Acta Cryst. B56, 697-714.]; Motherwell et al., 2002[Motherwell, W. D. S., Ammon, H. L., Dunitz, J. D., Dzyabchenko, A., Erk, P., Gavezzotti, A., Hofmann, D. W. M., Leusen, F. J. J., Lommerse, J. P. M., Mooij, W. T. M., Price, S. L., Scheraga, H., Schweizer, B., Schmidt, M. U., van Eijck, B. P., Verwer, P. & Williams, D. E. (2002). Acta Cryst. B58, 647-661.]; Day et al., 2005[Day, G. M., Motherwell, W. D. S., Ammon, H. L., Boerrigter, S. X. M., Della Valle, R. G., Venuti, E., Dzyabchenko, A., Dunitz, J. D., Schweizer, B., van Eijck, B. P., Erk, P., Facelli, J. C., Bazterra, V. E., Ferraro, M. B., Hofmann, D. W. M., Leusen, F. J. J., Liang, C., Pantelides, C. C., Karamertzanis, P. G., Price, S. L., Lewis, T. C., Nowell, H., Torrisi, A., Scheraga, H. A., Arnautova, Y. A., Schmidt, M. U. & Verwer, P. (2005). Acta Cryst. B61, 511-527.], 2009[Day, G. M., Cooper, T. G., Cruz-Cabeza, A. J., Hejczyk, K. E., Ammon, H. L., Boerrigter, S. X. M., Tan, J. S., Della Valle, R. G., Venuti, E., Jose, J., Gadre, S. R., Desiraju, G. R., Thakur, T. S., van Eijck, B. P., Facelli, J. C., Bazterra, V. E., Ferraro, M. B., Hofmann, D. W. M., Neumann, M. A., Leusen, F. J. J., Kendrick, J., Price, S. L., Misquitta, A. J., Karamertzanis, P. G., Welch, G. W. A., Scheraga, H. A., Arnautova, Y. A., Schmidt, M. U., van de Streek, J., Wolf, A. K. & Schweizer, B. (2009). Acta Cryst. B65, 107-125.]; Bardwell et al., 2011[Bardwell, D. A., Adjiman, C. S., Arnautova, Y. A., Bartashevich, E., Boerrigter, S. X. M., Braun, D. E., Cruz-Cabeza, A. J., Day, G. M., Della Valle, R. G., Desiraju, G. R., van Eijck, B. P., Facelli, J. C., Ferraro, M. B., Grillo, D., Habgood, M., Hofmann, D. W. M., Hofmann, F., Jose, K. V. J., Karamertzanis, P. G., Kazantsev, A. V., Kendrick, J., Kuleshova, L. N., Leusen, F. J. J., Maleev, A. V., Misquitta, A. J., Mohamed, S., Needs, R. J., Neumann, M. A., Nikylov, D., Orendt, A. M., Pal, R., Pantelides, C. C., Pickard, C. J., Price, L. S., Price, S. L., Scheraga, H. A., van de Streek, J., Thakur, T. S., Tiwari, S., Venuti, E. & Zhitkov, I. K. (2011). Acta Cryst. B67, 535-551.]; Reilly et al., 2016[Reilly, A. M., Cooper, R. I., Adjiman, C. S., Bhattacharya, S., Boese, A. D., Brandenburg, J. G., Bygrave, P. J., Bylsma, R., Campbell, J. E., Car, R., Case, D. H., Chadha, R., Cole, J. C., Cosburn, K., Cuppen, H. M., Curtis, F., Day, G. M., DiStasio, R. A. Jr, Dzyabchenko, A., van Eijck, B. P., Elking, D. M., van den Ende, J. A., Facelli, J. C., Ferraro, M. B., Fusti-Molnar, L., Gatsiou, C.-A., Gee, T. S., de Gelder, R., Ghiringhelli, L. M., Goto, H., Grimme, S., Guo, R., Hofmann, D. W. M., Hoja, J., Hylton, R. K., Iuzzolino, L., Jankiewicz, W., de Jong, D. T., Kendrick, J., de Klerk, N. J. J., Ko, H.-Y., Kuleshova, L. N., Li, X., Lohani, S., Leusen, F. J. J., Lund, A. M., Lv, J., Ma, Y., Marom, N., Masunov, A. E., McCabe, P., McMahon, D. P., Meekes, H., Metz, M. P., Misquitta, A. J., Mohamed, S., Monserrat, B., Needs, R. J., Neumann, M. A., Nyman, J., Obata, S., Oberhofer, H., Oganov, A. R., Orendt, A. M., Pagola, G. I., Pantelides, C. C., Pickard, C. J., Podeszwa, R., Price, L. S., Price, S. L., Pulido, A., Read, M. G., Reuter, K., Schneider, E., Schober, C., Shields, G. P., Singh, P., Sugden, I. J., Szalewicz, K., Taylor, C. R., Tkatchenko, A., Tuckerman, M. E., Vacarro, F., Vasileiadis, M., Vazquez-Mayagoitia, A., Vogt, L., Wang, Y., Watson, R. E., de Wijs, G. A., Yang, J., Zhu, Q. & Groom, C. R. (2016). Acta Cryst. B72, 439-459.]; Hunnisett et al., 2024[Hunnisett, L. M., Francia, N., Nyman, J., Abraham, N. S., Aitipamula, S., Alkhidir, T., Almehairbi, M., Anelli, A., Anstine, D. M., Anthony, J. E., Arnold, J. E., Bahrami, F., Bellucci, M. A., Beran, G. J. O., Bhardwaj, R. M., Bianco, R., Bis, J. A., Boese, A. D., Bramley, J., Braun, D. E., Butler, P. W. V., Cadden, J., Carino, S., Červinka, C., Chan, E. J., Chang, C., Clarke, S. M., Coles, S. J., Cook, C. J., Cooper, R. I., Darden, T., Day, G. M., Deng, W., Dietrich, H., DiPasquale, A., Dhokale, B., van Eijck, B. P., Elsegood, M. R. J., Firaha, D., Fu, W., Fukuzawa, K., Galanakis, N., Goto, H., Greenwell, C., Guo, R., Harter, J., Helfferich, J., Hoja, J., Hone, J., Hong, R., Hušák, M., Ikabata, Y., Isayev, O., Ishaque, O., Jain, V., Jin, Y., Jing, A., Johnson, E. R., Jones, I., Jose, K. V. J., Kabova, E. A., Keates, A., Kelly, P. F., Klimeš, J., Kostková, V., Li, H., Lin, X., List, A., Liu, C., Liu, Y. M., Liu, Z., Lončarić, I., Lubach, J. W., Ludík, J., Marom, N., Matsui, H., Mattei, A., Mayo, R. A., Melkumov, J. W., Mladineo, B., Mohamed, S., Momenzadeh Abardeh, Z., Muddana, H. S., Nakayama, N., Nayal, K. S., Neumann, M. A., Nikhar, R., Obata, S., O'Connor, D., Oganov, A. R., Okuwaki, K., Otero-de-la-Roza, A., Parkin, S., Parunov, A., Podeszwa, R., Price, A. J. A., Price, L. S., Price, S. L., Probert, M. R., Pulido, A., Ramteke, G. R., Rehman, A. U., Reutzel-Edens, S. M., Rogal, J., Ross, M. J., Rumson, A. F., Sadiq, G., Saeed, Z. M., Salimi, A., Sasikumar, K., Sekharan, S., Shankland, K., Shi, B., Shi, X., Shinohara, K., Skillman, A. G., Song, H., Strasser, N., van de Streek, J., Sugden, I. J., Sun, G., Szalewicz, K., Tan, L., Tang, K., Tarczynski, F., Taylor, C. R., Tkatchenko, A., Tom, R., Touš, P., Tuckerman, M. E., Unzueta, P. A., Utsumi, Y., Vogt-Maranto, L., Weatherston, J., Wilkinson, L. J., Willacy, R. D., Wojtas, L., Woollam, G. R., Yang, Y., Yang, Z., Yonemochi, E., Yue, X., Zeng, Q., Zhou, T., Zhou, Y., Zubatyuk, R. & Cole, J. C. (2024). Acta Cryst. B80, 548-574.]).

Results from the blind tests and other studies demonstrate that reliable CSP for rigid organic molecules is quickly becoming routine (Whittleton et al., 2017a[Whittleton, S. R., Otero-de-la-Roza, A. & Johnson, E. R. (2017a). J. Chem. Theory Comput. 13, 441-450.]; Price et al., 2023[Price, A. J., Mayo, R. A., Otero-de-la-Roza, A. & Johnson, E. R. (2023). CrystEngComm, 25, 953-960.]; Nikhar & Szalewicz, 2022[Nikhar, R. & Szalewicz, K. (2022). Nat. Commun. 13, 3095.]). While crystal structure predictions for flexible molecules are frequently successful as well (Neumann et al., 2015[Neumann, M. A., van de Streek, J., Fabbiani, F. P. A., Hidber, P. & Grassmann, O. (2015). Nat. Commun. 6, 7793. ]; Whittleton et al., 2017b[Whittleton, S. R., Otero-de-la-Roza, A. & Johnson, E. R. (2017b). J. Chem. Theory Comput. 13, 5332-5342.]; Braun et al., 2017[Braun, D. E., Lingireddy, S. R., Beidelschies, M. D., Guo, R., Müller, P., Price, S. L. & Reutzel-Edens, S. M. (2017). Cryst. Growth Des. 17, 5349-5365.], 2019[Braun, D. E., McMahon, J. A., Bhardwaj, R. M., Nyman, J., Neumann, M. A., van de Streek, J. & Reutzel-Edens, S. M. (2019). Cryst. Growth Des. 19, 2947-2962.]; Mortazavi et al., 2019[Mortazavi, M., Hoja, J., Aerts, L., Quéré, L., van de Streek, J., Neumann, M. A. & Tkatchenko, A. (2019). Commun. Chem. 2, 70.]; Hoja et al., 2019[Hoja, J., Ko, H., Neumann, M. A., Car, R., DiStasio, R. A. Jr & Tkatchenko, A. (2019). Sci. Adv. 5, eaau3338.]; Bhardwaj et al., 2019[Bhardwaj, R. M., McMahon, J. A., Nyman, J., Price, L. S., Konar, S., Oswald, I. D. H., Pulham, C. R., Price, S. L. & Reutzel-Edens, S. M. (2019). J. Am. Chem. Soc. 141, 13887-13897.]; Firaha et al., 2023[Firaha, D., Liu, Y. M., van de Streek, J., Sasikumar, K., Dietrich, H., Helfferich, J., Aerts, L., Braun, D. E., Broo, A., DiPasquale, A. G., Lee, A. Y., Le Meur, S., Nilsson Lill, S. O., Lunsmann, W. J., Mattei, A., Muglia, P., Putra, O. D., Raoui, M., Reutzel-Edens, S. M., Rome, S., Sheikh, A. Y., Tkatchenko, A., Woollam, G. R. & Neumann, M. A. (2023). Nature, 623, 324-328.]), the prediction of conformational polymorphs remains more challenging than rigid-molecule CSP for multiple reasons. First, having numerous degrees of intramolecular conformational freedom dramatically increases the search space of potential crystal structures (Oganov, 2018[Oganov, A. R. (2018). Faraday Discuss. 211, 643-660.]). Second, the selection of conformations to consider for solid-state packing is complicated by the fact that gas-phase conformational energies can be a poor proxy for crystalline stability (Thompson & Day, 2014[Thompson, H. P. G. & Day, G. M. (2014). Chem. Sci. 5, 3173-3182.]). For example, whereas an extended conformation of a molecule might be considerably less stable than a folded one in the gas phase, intermolecular interactions in the solid state may preferentially stabilize the extended conformation. Thus, a CSP search may need to consider alternate criteria beyond gas-phase conformational energies when generating crystal structures.

Third, capturing the interplay of intra- and intermolecular interactions properly in the crystal structure optimization and energy ranking has been a long-standing challenge in CSP. Many years ago, it was recognized that typical classical mechanical force fields frequently lack the requisite accuracy for intramolecular conformational energies (Karamertzanis et al., 2008[Karamertzanis, P. G., Day, G. M., Welch, G. W. A., Kendrick, J., Leusen, F. J. J., Neumann, M. A. & Price, S. L. (2008). J. Chem. Phys. 128, 244708.]; Price, 2008[Price, S. L. (2008). Int. Rev. Phys. Chem. 27, 541-568.]). This motivated the development of models which combined quantum mechanical descriptions of the intramolecular energy (either computed directly or via custom-fitted potentials) with force field intermolecular energies (Bowskill et al., 2021[Bowskill, D. H., Sugden, I. J., Konstantinopoulos, S., Adjiman, C. S. & Pantelides, C. C. (2021). Annu. Rev. Chem. Biomol. Eng. 12, 593-623.]).

The widespread successes of van der Waals-inclusive density functional theory (DFT) methods in the fourth blind test (Neumann et al., 2008[Neumann, M. A., Leusen, F. J. J. & Kendrick, J. (2008). Angew. Chem. Int. Ed. 47, 2427-2430.]) and many subsequent studies have shifted many researchers in the field away from such hybrid intra-/intermolecular approaches, at least in the final stage(s) of CSP structure refinement and ranking. However, despite many successful structure predictions, increasing numbers of examples have been found in the past several years where widely-used generalized-gradient approximation (GGA) and hybrid density functionals (particularly those with ∼20–25% exact exchange) perform surprisingly poorly. For some acid–base co-crystals, GGA functionals spuriously transfer protons and incorrectly predict a salt form to be more stable than the neutral co-crystal (LeBlanc et al., 2018[LeBlanc, L. M., Dale, S. G., Taylor, C. R., Becke, A. D., Day, G. M. & Johnson, E. R. (2018). Angew. Chem. Int. Ed. 57, 14906-14910.]). In the molecule ROY, named for its red, orange, and yellow polymorphs, a variety of DFT methods erroneously predict form Y to be among the least stable polymorphs, when it is actually the most stable experimentally (Tan et al., 2018[Tan, M., Shtukenberg, A. G., Zhu, S., Xu, W., Dooryhee, E., Nichols, S. M., Ward, M. D., Kahr, B. & Zhu, Q. (2018). Faraday Discuss. 211, 477-491.]; Nyman et al., 2019[Nyman, J., Yu, L. & Reutzel-Edens, S. M. (2019). CrystEngComm, 21, 2080-2088.]; Greenwell et al., 2020[Greenwell, C., McKinley, J. L., Zhang, P., Zeng, Q., Sun, G., Li, B., Wen, S. & Beran, G. J. O. (2020). Chem. Sci. 11, 2200-2214.]).

The common origin of these poor predictions lies in density-driven delocalization error in the approximate density functionals. In the acid–base co-crystals, GGAs over-stabilize the delocalized charges of the salt forms. In ROY, delocalization error causes GGAs and many hybrids to over-stabilize more planar molecular conformations with extended π-conjugation, such as those found in the red and orange polymorphs, over the nearly perpendicular conformations with more localized electron densities that occur in the yellow polymorphs (Thomas & Spackman, 2018[Thomas, S. P. & Spackman, M. A. (2018). Aust. J. Chem. 71, 279.]; Nyman et al., 2019[Nyman, J., Yu, L. & Reutzel-Edens, S. M. (2019). CrystEngComm, 21, 2080-2088.]; Beran et al., 2022a[Beran, G. J. O., Sugden, I. J., Greenwell, C., Bowskill, D. H., Pantelides, C. C. & Adjiman, C. S. (2022a). Chem. Sci. 13, 1288-1297.]). Beyond ROY, conformational energy errors arising from delocalization error affect crystal polymorph rankings in pharmaceuticals, organic semi-conductor materials, molecule X from the third blind test, and other small molecules (Whittleton et al., 2017b[Whittleton, S. R., Otero-de-la-Roza, A. & Johnson, E. R. (2017b). J. Chem. Theory Comput. 13, 5332-5342.]; Greenwell & Beran, 2020[Greenwell, C. & Beran, G. J. O. (2020). Cryst. Growth Des. 20, 4875-4881.], 2021[Greenwell, C. & Beran, G. J. O. (2021). J. Mater. Chem. C, 9, 2848-2857.]). They also impact the chemical reaction energies in photochromic materials (Beran, 2019[Beran, G. J. O. (2019). CrystEngComm, 21, 758-764.]; Gately et al., 2021[Gately, T. J., Sontising, W., Easley, C. J., Islam, I., Al-Kaysi, R. O., Beran, G. J. O. & Bardeen, C. J. (2021). CrystEngComm, 23, 5931-5943.], 2022[Gately, T. J., Cook, C. C., Almuzarie, R., Islam, I., Gardner, Z. T., Iuliucci, R. J., Al-Kaysi, R. O., Beran, G. J. O. & Bardeen, C. J. (2022). Cryst. Growth Des. 22, 7298-7307.]; Cook et al., 2022[Cook, C. J., Li, W., Lui, B. F., Gately, T. J., Al-Kaysi, R. O., Mueller, L. J., Bardeen, C. J. & Beran, G. J. O. (2023). Chem. Sci. 14, 937-949.], 2023[Cook, C. J., Perry, C. J. & Beran, G. J. O. (2023). J. Phys. Chem. Lett. 14, 6823-6831.]; Perry & Beran, 2023[Perry, C. J. & Beran, G. J. O. (2023). Cryst. Growth Des. 23, 8352-8360.]) and the lattice energies in halogen-bonded crystals (Otero-de-la Roza et al., 2019[Otero-de-la-Roza, A., LeBlanc, L. M. & Johnson, E. R. (2019). J. Chem. Theory Comput. 15, 4933-4944.]).

Delocalization error issues can potentially be addressed using hybrid functionals that include a larger fraction (e.g. ∼50%) of exact exchange (Price et al., 2022[Price, A. J., Otero-de-la-Roza, A. & Johnson, E. R. (2022). Chem. Sci. 14, 1252-1262.], 2023[Price, A. J., Mayo, R. A., Otero-de-la-Roza, A. & Johnson, E. R. (2023). CrystEngComm, 25, 953-960.]), though care is needed to ensure the density functional remains well-balanced overall despite the larger exact exchange contribution. For example, BHLYP-D3(BJ), which includes 50% exchange, performs notably worse than B3LYP-D3(BJ) for hydrogen-bonding interactions (Řezáč, 2020a[Řezáč, J. (2020a). J. Chem. Theory Comput. 16, 2355-2368.], Řezáč, 2020b[Řezáč, J. (2020b). J. Chem. Theory Comput. 16, 6305-6316.]). For the cases where the delocalization error is primarily intramolecular in nature, another strategy returns to the idea of decoupling the intra- and intermolecular interactions by combining crystalline periodic DFT lattice energies with a molecular conformational energy correction that is evaluated for the isolated molecule(s) taken from the crystal (Greenwell & Beran, 2020[Greenwell, C. & Beran, G. J. O. (2020). Cryst. Growth Des. 20, 4875-4881.]). Because gas-phase calculations on an isolated molecule are relatively inexpensive, the intramolecular correction can be evaluated with more advanced/robust electronic structure models such as the range-separated hybrid functional ωB97M-V (Mardirossian & Head-Gordon, 2016[Mardirossian, N. & Head-Gordon, M. (2016). J. Chem. Phys. 144, 214110.]), double-hybrid functionals such as revDSD-PBEP86-D4 (Santra et al., 2019[Santra, G., Sylvetsky, N. & Martin, J. M. L. (2019). J. Phys. Chem. A, 123, 5129-5143.]), spin-component-scaled dispersion-corrected second-order Møller–Plesset perturbation theory (SCS-MP2D) (Greenwell et al., 2022[Greenwell, C., Řezáč, J. & Beran, G. J. O. (2022). Phys. Chem. Chem. Phys. 24, 3695-3712.]), or domain-local pair-natural orbital coupled cluster singles, doubles, and perturbative triples (DLPNO-CCSD(T)) (Riplinger et al., 2013[Riplinger, C., Sandhoefer, B., Hansen, A. & Neese, F. (2013). J. Chem. Phys. 139, 134101.], 2016[Riplinger, C., Pinski, P., Becker, U., Valeev, E. F. & Neese, F. (2016). J. Chem. Phys. 144, 024109.]).

When applied to the crystal structure prediction for ROY, for example, the combination of dispersion-corrected periodic DFT lattice energies (B86bPBE-XDM) and intramolecular SCS-MP2D corrections produced the first crystal energy landscape that was generally consistent with experimentally-known polymorph stabilities (Beran et al., 2022a[Beran, G. J. O., Sugden, I. J., Greenwell, C., Bowskill, D. H., Pantelides, C. C. & Adjiman, C. S. (2022a). Chem. Sci. 13, 1288-1297.]). The same approach improves the polymorph rankings for systems such as molecule X or the pharmaceuticals axitinib and galunisertib (Greenwell & Beran, 2020[Greenwell, C. & Beran, G. J. O. (2020). Cryst. Growth Des. 20, 4875-4881.]), and it produces much more sensible reaction energetics for the solid-state photochromic reactions described above. Several examples where intramolecular corrections have proved useful have been reviewed recently (Beran et al., 2023[Beran, G. J. O., Greenwell, C., Cook, C. & Řezáč, J. (2023). Acc. Chem. Res. 56, 3525-3534.]), and such intramolecular corrections have been incorporated in commercial CSP work flows (Firaha et al., 2023[Firaha, D., Liu, Y. M., van de Streek, J., Sasikumar, K., Dietrich, H., Helfferich, J., Aerts, L., Braun, D. E., Broo, A., DiPasquale, A. G., Lee, A. Y., Le Meur, S., Nilsson Lill, S. O., Lunsmann, W. J., Mattei, A., Muglia, P., Putra, O. D., Raoui, M., Reutzel-Edens, S. M., Rome, S., Sheikh, A. Y., Tkatchenko, A., Woollam, G. R. & Neumann, M. A. (2023). Nature, 623, 324-328.]).

Of course, these intramolecular corrections address only one manifestation of delocalization error, and intermolecular delocalization error can still be problematic in some systems (such as the acid–base co-crystals mentioned above). Cases have been found where there are significant errors in both the intra- and intermolecular interactions that fortuitously cancel each other somewhat in the periodic DFT treatment (Beran et al., 2022b[Beran, G. J. O., Wright, S. E., Greenwell, C. & Cruz-Cabeza, A. J. (2022b). J. Chem. Phys. 156, 104112.]). In those systems, correcting the intramolecular error can have the unintended effect of exposing the intermolecular errors more fully, leading to worse agreement with experiment. On the other hand, relying on the cancellation of intra- and intermolecular errors can also be problematic, since the extent of error cancellation may differ across different polymorphs on the crystal energy landscape.

Given the clear challenges associated with modeling conformational polymorphs, the present study investigates the impact of conformational energies on the polymorph energy landscapes of molecules XXXI and XXXII (Fig. 1[link]) from the seventh blind test. Molecule XXXII exhibits many flexible dihedral angles, including several that might impact the extent of π-conjugation and make its polymorph energy rankings susceptible to density-driven delocalization error with some DFT functionals. In contrast, while molecule XXXI exhibits considerable flexibility, that flexibility does not alter the extent of the π-conjugation, reducing the likely impact of intramolecular delocalization error. Thus, these two systems provide an interesting comparison of how DFT and other methods perform for conformational polymorphs when π-conjugation-driven delocalization error either is or is not prominent.

[Figure 1]
Figure 1
Molecules XXXI and XXXII from the seventh blind test of CSP. Arrows indicate flexible dihedral angles. For molecule XXXII, rotation about dihedral angles d1d4 in red impacts the extent of π-conjugation.

This study originated with our participation in the structure ranking phase (Phase 2) of the seventh blind test, but the investigation here has been extended beyond the results we submitted for the test. In the sections that follow, we investigate the crystal energy landscapes for both species with and without conformational energy corrections, examine the nature of the conformational energy errors, and compare the performance of a variety of semi-empirical, DFT, and correlated wavefunction methods for describing those conformational energies. Notably, while modeling the conformational energies of molecule XXXI proves relatively straightforward, our results highlight the substantial difficulties that remain when trying to study highly flexible species such as molecule XXXII.

2. Computational methods

During phase 2 of the seventh blind test, the Cambridge Crystallographic Data Centre (CCDC) provided 100 candidate structures of molecule XXXI and 500 candidate structures of molecule XXXII. The methods we used to select, refine, and rank the crystal structures from these lists are described below. To refer to structures from these sets, we adopt the nomenclature that, for example, #055 refers to structure 55 from original structure list provided for that species by the CCDC. The final DFT-optimized crystal structures used in this study are provided in supporting information.

2.1. Crystal structure selection and geometry optimizations

Molecule XXXI. Ninety-nine of the 100 crystal structures provided during Phase 2 of the blind test were fully relaxed (both atomic positions and unit-cell parameters) with periodic DFT. The remaining structure, #089 (Z = 18), was omitted for computational expedience, though this structure was later revealed to be experimental form C. The DFT calculations employed the B86bPBE density functional (Becke, 1986[Becke, A. D. (1986). J. Chem. Phys. 85, 7184-7187.]; Perdew et al., 1996[Perdew, J. P., Burke, K. & Ernzerhof, M. (1996). Phys. Rev. Lett. 77, 3865-3868.]) and exchange-hole dipole moment (XDM) dispersion correction (Otero-de-la Roza & Johnson, 2012[Otero-de-la-Roza, A. & Johnson, E. R. (2012). J. Chem. Phys. 136, 174109.]). The projector augmented wave (PAW) approach, a 50 Ry planewave cutoff and Monkhorst–Pack k-point grid spacing of at least 0.06 Å−1 were used. The variable-cell QuantumEspresso (Giannozzi et al., 2017[Giannozzi, P., Andreussi, O., Brumme, T., Bunau, O., Buongiorno Nardelli, M., Calandra, M., Car, R., Cavazzoni, C., Ceresoli, D., Cococcioni, M., Colonna, N., Carnimeo, I., Dal Corso, A., de Gironcoli, S., Delugas, P., DiStasio, R. A., Ferretti, A., Floris, A., Fratesi, G., Fugallo, G., Gebauer, R., Gerstmann, U., Giustino, F., Gorni, T., Jia, J., Kawamura, M., Ko, H.-Y., Kokalj, A., Küçükbenli, E., Lazzeri, M., Marsili, M., Marzari, N., Mauri, F., Nguyen, N. L., Nguyen, H.-V., Otero-de-la-Roza, A., Paulatto, L., Poncé, S., Rocca, D., Sabatini, R., Santra, B., Schlipf, M., Seitsonen, A. P., Smogunov, A., Timrov, I., Thonhauser, T., Umari, P., Vast, N., Wu, X. & Baroni, S. (2017). J. Phys. Condens. Matter, 29, 465901.]) crystal structure optimizations employed energy and geometry convergence criteria of etot_conv_thr = 2 × 10−6 a.u. and forc_conv_thr = 6 × 10−4 a.u.

Molecule XXXII. Given the large number of crystal structures and the large unit cell sizes for molecule XXXII, a hierarchical refinement was used to select structures for full DFT optimization and ranking. The 500 provided structures were initially optimized with the semi-empirical HF-3c model (Sure & Grimme, 2013[Sure, R. & Grimme, S. (2013). J. Comput. Chem. 34, 1672-1685.]) under periodic boundary conditions using Crystal17 (Dovesi et al., 2018[Dovesi, R., Erba, A., Orlando, R., Zicovich-Wilson, C. M., Civalleri, B., Maschio, L., Rérat, M., Casassa, S., Baima, J., Salustro, S. & Kirtman, B. (2018). WIREs Comput. Mol. Sci. 8, e1360.]). The original solid-state implementation of HF-3c was used (Brandenburg & Grimme, 2014[Brandenburg, J. G. & Grimme, S. (2014). In Prediction and Calculation of Crystal Structures: Methods and Applications, Vol. 345, Topics in Current Chemistry, edited by S. Atahan-Evrenk & A. Aspuru-Guzik, pp. 1-24. Cham: Springer International Publishing.]), rather than the rescaled s-HF-3c variant (Cutini et al., 2016[Cutini, M., Civalleri, B., Corno, M., Orlando, R., Brandenburg, J. G., Maschio, L. & Ugliengo, P. (2016). J. Chem. Theory Comput. 12, 3340-3352.]). When it became apparent during the course of the blind test that those structures differed too strongly from the final DFT ones to provide a reliable preliminary ranking, 481 of the structures were then loosely optimized with periodic DFT, again using the B86bPBE-XDM functional. These preliminary DFT optimizations employed slightly less dense k-point grids and looser convergence criteria (etot_conv_thr = 10−4 a.u. and forc_conv_thr = 10−3 a.u). The remaining 19 structures were omitted due to their larger unit cells and comparatively high HF-3c energies. During the blind test, all 33 low-energy structures lying within ∼5 kJ mol−1 of the global minimum energy (based either on the DFT energies or the SCS-MP2D conformational energy-corrected single-point energies described below) were optimized more tightly with the same settings applied to molecule XXXI. This 5 kJ mol−1 cutoff threshold was chosen due to practical time constraints during the test, but it unfortunately excluded the experimental structures.

After the blind test results were released, the set of tightly-refined crystal structures was enlarged for the present study to include the 50 most stable structures on the list we had submitted to the blind test, which corresponded to all structures within 7.9 kJ mol−1 of the global minimum (GM) at the level of B86bPBE-XDM with SCS-MP2D conformational energy corrections. This expanded list includes a total of 53 structures from the initial 500, and it includes the experimental forms Amaj and B.

2.2. Single-point energies

To address delocalization error or other limitations in the B86bPBE-XDM conformational energies, final single-point energy calculations were performed that correct the intramolecular conformational energies with a higher level of theory (Greenwell & Beran, 2020[Greenwell, C. & Beran, G. J. O. (2020). Cryst. Growth Des. 20, 4875-4881.])

[E_{\rm crystal}^{\rm corrected}=E_{\rm crystal}^{\rm DFT}+\sum_{i}^{Z}\left(E_ {{\rm mon},i}^{\rm higher}-E_{{\rm mon},i}^{\rm DFT}\right).\eqno(1)]

In this expression, the periodic DFT energy of the crystal ([E_{\rm crystal}^{\rm DFT}]) is corrected via gas-phase monomer calculations performed with the same DFT model ([E_{{\rm mon},i}^{\rm DFT}]) and at a higher level of theory ([E_{{\rm mon},i}^{\rm higher}]). The monomer geometries are extracted directly from each DFT-optimized crystal. The sum in equation (1[link]) runs over all monomers in the unit cell, though exploitation of space-group symmetry reduces the number of unique monomers that need to be computed to one (if Z′ = 1) or two (if Z′ = 2). The higher levels of theory used here include DLPNO-CCSD(T1) (Guo et al., 2018[Guo, Y., Riplinger, C., Becker, U., Liakos, D. G., Minenkov, Y., Cavallo, L. & Neese, F. (2018). J. Chem. Phys. 148, 011101.]), spin-component-scaled dispersion-corrected second-order Møller–Plesset perturbation theory (SCS-MP2D), and different density functionals. Further details of the gas-phase conformational energy calculations are discussed below.

2.3. Gas-phase conformational energies

Because a few of the crystals in the molecule XXXI and XXXII sets have Z′ = 2, there are a total of 100 symmetrically unique conformations of molecule XXXI and 55 conformations of molecule XXXII. The B86bPBE-XDM gas-phase conformational energies in equation (1[link]) were computed in QuantumEspresso, using the same 50 Ry planewave cutoff and PAW potentials as for the crystal calculations. The molecules were placed in a large orthorhombic unit cell that ensures separation between any atoms in the molecule and its periodic images of at least 20 Å in all directions, and only the Γ-point was sampled.

SCS-MP2D conformational energies were obtained by first obtaining MP2 energies that were extrapolated to the complete basis set (CBS) limit from aug-cc-pVTZ and aug-cc-pVQZ (Dunning, 1989[Dunning, T. H. (1989). J. Chem. Phys. 90, 1007-1023.]) results by combining Hartree–Fock (HF)/aug-cc-pVQZ with CBS-limit correlation energies according to Helgaker et al. (1997[Helgaker, T., Klopper, W., Koch, H. & Noga, J. (1997). J. Chem. Phys. 106, 9639-9646.]),

[E_{\rm CBS}^{\rm MP2}=E_{\rm aQZ}^{\rm HF}+E_{\rm CBS}^{\rm corr}=E_{\rm aQZ}^ {\rm HF}+{{4^{3}E^{\rm corr}_{\rm aQZ}-3^{3}E^{\rm corr}_{\rm aTZ}} \over {4^{3}- 3^{3}}}.\eqno(2)]

These MP2 calculations were performed with PSI4 (Smith et al., 2020[Smith, D. G. A., Burns, L. A., Simmonett, A. C., Parrish, R. M., Schieber, M. C., Galvelis, R., Kraus, P., Kruse, H., Di Remigio, R., Alenaizan, A., James, A. M., Lehtola, S., Misiewicz, J. P., Scheurer, M., Shaw, R. A., Schriber, J. B., Xie, Y., Glick, Z. L., Sirianni, D. A., O'Brien, J. S., Waldrop, J. M., Kumar, A., Hohenstein, E. G., Pritchard, B. P., Brooks, B. R., Schaefer, H. F., Sokolov, A. Y., Patkowski, K., DePrince, A. E., Bozkaya, U., King, R. A., Evangelista, F. A., Turney, J. M., Crawford, T. D. & Sherrill, C. D. (2020). J. Chem. Phys. 152, 184108.]). The final SCS-MP2D energies were obtained from the MP2 results using the MP2D library (Greenwell & Beran, 2018[Greenwell, C. & Beran, G. J. O. (2018). The MP2D software library. https://github.com/Chandemonium/MP2D.]).

DLPNO-CCSD(T1) energies were computed in Orca 5 (Neese, 2012[Neese, F. (2012). WIREs Comput. Mol. Sci. 2, 73-78.]) and extrapolated to the CBS limit via the focal point approach

[E_{\rm CBS}^{\rm DLPNO-CCSD(T1)}=E_{\rm CBS}^{\rm MP2}+E_{\rm aXZ}^{\rm DLPNO- CCSD(T1)}-E_{\rm aXZ}^{\rm MP2}, \eqno(3)]

where `aXZ' refers to a basis set in the aug-cc-pVXZ family (Dunning, 1989[Dunning, T. H. (1989). J. Chem. Phys. 90, 1007-1023.]). For molecule XXXI, the DLPNO-CCSD(T1) calculations use the aug-cc-pVTZ basis, TightPNO settings and TCutMKN = 10−4. The molecule XXXI results here differ slightly from those submitted to the blind test, where the non-iterative triples variant DLPNO-CCSD(T0) (Riplinger et al., 2013[Riplinger, C., Sandhoefer, B., Hansen, A. & Neese, F. (2013). J. Chem. Phys. 139, 134101.]) was used instead of the iterative triples correction in DLPNO-CCSD(T1). The mean absolute deviation in the relative conformational energies between T1 and T0 triples across all molecule XXXI structures is only 0.04 kJ mol−1, with a maximum deviation of 0.19 kJ mol−1. For further convergence testing, we examined the difference between using aug-cc-pVDZ and aug-cc-pVTZ energies in the CBS-limit extrapolation of the DLPNO-CCSD(T) energies [equation (3)[link]]. At the DLPNO-CCSD(T0) level, the resulting conformational energies differed by only a mean absolute average of 0.10 kJ mol−1 (maximum 0.43 kJ mol−1). Similar deviations are found if one omits the diffuse basis functions by using cc-pVTZ instead of aug-cc-pVTZ. Finally, the impact of tightening the TCutPNO parameter by an additional factor of 3 (to 3.33 × 10−8) was also tested at the aug-cc-pVDZ level, and it altered the relative conformational energies only by a mean absolute average of 0.06 kJ mol−1 (maximum 0.15 kJ mol−1). Therefore, the default TightPNO setting of TCutPNO = 1.0 × 10−7 was used instead for the final aug-cc-pVTZ calculations. Details of the convergence tests can be found in Section S1.3 in supporting information. Taken together, these results suggest that the DLPNO-CCSD(T1) molecule XXXI conformational energies are likely converged to within a few tenths of a kJ mol−1 or better.

Converging the DLPNO-CCSD(T) calculations for molecule XXXII proves more difficult. First, due to the large molecular size, the aug-cc-pVDZ basis set (1585 basis functions) was used instead of aug-cc-pVTZ (3047 basis functions). For molecule XXXI, the convergence testing above found only modest differences between the two basis sets once extrapolated to the CBS limit. However, the discrepancies might be larger for molecule XXXII, due to its greater complexity and the more variable magnitudes of the intra-molecular non-covalent interactions across the different conformations. For example, the cc-pVDZ and aug-cc-pVDZ conformational energies of molecule XXXII vary by a mean absolute deviation of 0.39 kJ mol−1, compared to 0.16 kJ mol−1 for molecule XXXI with the same basis sets. If one uses molecule XXXI as a guide, this would imply that similarly large differences would be found if one used aug-cc-pVTZ instead of aug-cc-pVDZ. Second, the relative conformational energies in molecule XXXII are more sensitive to the numerical thresholds and triples treatment. Convergence testing found that tightening TCutPNO from 10−7 to 3.33 × 10−8 and employing the iterative T1 triples instead of non-iterative T0 altered the relative conformational energies by mean absolute 0.28 kJ mol−1 on average. While many of the conformational energies changed only ∼0.1–0.2 kJ mol−1, the conformational energies for several highly-folded conformers changed by a much larger 1.0–1.3 kJ mol−1 (with sizable contributions arising from both tightening TCutPNO and using the iterative T1 triples correction). See supporting information (Section 2.4) for details. Accordingly, the DLPNO-CCSD(T1) with the tighter TCutPNO setting was used for the final benchmark conformational energies here. Overall, given that the sensitivity of the benchmark XXXII conformational energies to the basis set and DLPNO parameters are a few times larger than for XXXI, we coarsely estimate that the molecule XXXII conformational energy uncertainties are probably at least ∼0.5 kJ mol−1, and they may be larger for the highly folded conformations.

Gas-phase conformational energies have also been evaluated with several additional DFT and semi-empirical models: GGA functional PBE-D4 (Perdew et al., 1996[Perdew, J. P., Burke, K. & Ernzerhof, M. (1996). Phys. Rev. Lett. 77, 3865-3868.]; Caldeweyher et al., 2017[Caldeweyher, E., Bannwarth, C. & Grimme, S. (2017). J. Chem. Phys. 147, 034112.]), hybrid functional PBE0-D4 (Adamo & Barone, 1999[Adamo, C. & Barone, V. (1999). J. Chem. Phys. 110, 6158-6170.]), range-separated hybrid meta-GGA ωB97M-V, and double-hybrid functional revDSD-PBEP86-D4. PBE and PBE0 are included because of their widespread use in crystal structure prediction, while ωB97M-V and revDSD-PBEP86-D4 are representative of the best-performing density functionals from large-scale benchmark studies (Martin & Santra, 2020[Martin, J. M. L. & Santra, G. (2020). Isr. J. Chem. 60, 787-804.]; Řezáč, 2022[Řezáč, J. (2022). Phys. Chem. Chem. Phys. 24, 14780-14793.]). The PBE-D4, PBE0-D4, and ωB97M-V calculations were performed with PSI4, while the remaining methods were computed using Orca. Most of these DFT calculations were performed in the aug-cc-pVQZ basis set. However, the revDSD-PBEP86-D4 calculations for molecule XXXII employed def2-QZVP instead due to frequent issues in converging the self-consistent-field equations with the larger aug-cc-pVQZ basis. Test calculations on molecule XXXI found that the revDSD-PBEP86-D4 root-mean-square errors relative to DLPNO-CCSD(T1) differed by only 0.02 kJ mol−1 between def2-QZVP and aug-cc-pVQZ, suggesting that the impact of using the smaller basis set for molecule XXXII is also probably small.

Three inexpensive semi-empirical models are also considered: HF-3c, PBEh-3c (Grimme et al., 2015[Grimme, S., Brandenburg, J. G., Bannwarth, C. & Hansen, A. (2015). J. Chem. Phys. 143, 054107.]), and r2SCAN-3c (Grimme et al., 2021[Grimme, S., Hansen, A., Ehlert, S. & Mewes, J.-M. (2021). J. Chem. Phys. 154, 064103.]). These were chosen because of their potential use for intermediate-level refinement/ranking of candidate crystal structures. These calculations were performed with Orca.

Finally, the sections below analyze the errors in the conformational energies. It is important to recognize that the error statistics obtained when comparing two methods can differ (sometimes significantly) depending on the choice of the reference conformer. The lowest-energy conformer is commonly chosen for the reference conformation. For molecule XXXII, however, it was found that a number of models disagree significantly with DLPNO-CCSD(T1) on the stability of this most stable conformer, and using this structure as the reference energy effectively imparts this disagreement into the relative conformational energies of all other conformations. To reduce the biases introduced by selecting any one particular conformation, the gas-phase conformational energies discussed below are computed relative to the average conformational energy computed at each level of theory. This choice of the reference energy effectively means that the conformational energy errors offer insight into the distribution of conformational energies about the average conformational energy.

2.4. Gas-phase conformational energy scans

To understand the energetics of the molecular conformations found in the crystals better, a series of one-dimensional (1-D) gas-phase conformational scans were performed about selected dihedral angles in each molecule. For each chosen value of the selected dihedral angle, all other degrees of freedom were relaxed using the same B86bPBE-XDM model as described above. Single-point energies with various models were then performed as described in Section 2.3[link].

For molecule XXXI, a scan was performed about dihedral angle d2 as defined in Fig. 1[link]. For molecule XXXII, 1-D scans were performed about each of d1d4. For computational expedience, the molecule XXXII scans were performed on fragments of XXXII instead of the whole molecule. The fragment used for each scan is shown in the corresponding potential energy curve figure, and any truncated bonds were terminated with hydrogen atoms.

3. Results and discussion

3.1. Molecule XXXI

Three polymorphs of XXXI have been found experimentally. Forms A and B are related enantiotropically, with form B (#025, P21/c, Z = 4, Z′ = 1,) being the most stable form at lower temperatures, and form A (P21/c, Z = 4, Z′ = 1) becoming the thermodynamically preferred polymorph at higher temperatures (a transition occurs at ∼55°C). Form A has major and minor disorder components, which are represented via structures #098 (Amaj) and #001 (Amin), respectively. Form C (#089, [R\bar{3}], Z = 18, Z′ = 1) is solvent-templated and contains large void channels. Form C is the least stable form, at least when the pores are unoccupied, and it is omitted here due to its large unit-cell size (Section 2.1[link]).

Fig. 2[link](a) is a plot of the crystal energy landscape obtained after optimizing the 99 structures with periodic B86bPBE-XDM. This DFT functional predicts the major disorder component of form Amaj to be the global minimum (GM). Form B is the most stable experimentally, but B86bPBE-XDM predicts that it lies at rank 16, about 4 kJ mol−1 higher than Amaj. The minor component Amin lies at rank 4, only 0.8 kJ mol−1 above Amaj. Correcting the B86bPBE-XDM conformational energies with DLPNO-CCSD(T1) shifts the relative lattice energies modestly [Fig. 2[link](a)], and Fig. 2[link](b) is a plot of the final crystal energy landscape. Candidate structure #071 shifts from rank 3 at the B86bPBE-XDM level to become the new GM after the DLPNO-CCSD(T1) correction. The Amaj and Amin structures are destabilized and now sit at ranks 7 and 9, or 2.2–2.5 kJ mol−1 above the GM. Form B now lies 4.8 kJ mol−1 above the GM at rank 17. Though these conformational energy corrections shift the relative energies, they do not alter the qualitative stability ordering among the three experimental forms. The modest impact of the conformational energy corrections on the relative energies of the GM and experimental structures can be understood from the close similarities of the conformations found in all four crystal structures (Fig. 3[link]).

[Figure 2]
Figure 2
(a) Impact of applying intramolecular conformational energy corrections to the relative B86bPBE-XDM lattice energies of the candidate crystal structures for molecule XXXI. (b) Final B86bPBE-XDM + ΔDLPNO-CCSD(T1) crystal energy landscape for molecule XXXI. Form B is the most stable form experimentally. Tabulated energetics and additional CSP landscapes can be found in Section S1.1 of supporting information.
[Figure 3]
Figure 3
Crystal structures of the molecule XXXI experimental polymorphs, Amaj and B, and the predicted global minimum structure GM.

The conformational energy-corrected relative lattice energies still disagree with the experimental room-temperature stability ordering by placing Amaj and Amin below form B. No vibrational free energy calculations are performed here, though results reported by other blind test participants suggest that the inclusion of vibrational free contributions can preferentially stabilize form B relative to form A at room temperature (Hunnisett et al., 2024[Hunnisett, L. M., Francia, N., Nyman, J., Abraham, N. S., Aitipamula, S., Alkhidir, T., Almehairbi, M., Anelli, A., Anstine, D. M., Anthony, J. E., Arnold, J. E., Bahrami, F., Bellucci, M. A., Beran, G. J. O., Bhardwaj, R. M., Bianco, R., Bis, J. A., Boese, A. D., Bramley, J., Braun, D. E., Butler, P. W. V., Cadden, J., Carino, S., Červinka, C., Chan, E. J., Chang, C., Clarke, S. M., Coles, S. J., Cook, C. J., Cooper, R. I., Darden, T., Day, G. M., Deng, W., Dietrich, H., DiPasquale, A., Dhokale, B., van Eijck, B. P., Elsegood, M. R. J., Firaha, D., Fu, W., Fukuzawa, K., Galanakis, N., Goto, H., Greenwell, C., Guo, R., Harter, J., Helfferich, J., Hoja, J., Hone, J., Hong, R., Hušák, M., Ikabata, Y., Isayev, O., Ishaque, O., Jain, V., Jin, Y., Jing, A., Johnson, E. R., Jones, I., Jose, K. V. J., Kabova, E. A., Keates, A., Kelly, P. F., Klimeš, J., Kostková, V., Li, H., Lin, X., List, A., Liu, C., Liu, Y. M., Liu, Z., Lončarić, I., Lubach, J. W., Ludík, J., Marom, N., Matsui, H., Mattei, A., Mayo, R. A., Melkumov, J. W., Mladineo, B., Mohamed, S., Momenzadeh Abardeh, Z., Muddana, H. S., Nakayama, N., Nayal, K. S., Neumann, M. A., Nikhar, R., Obata, S., O'Connor, D., Oganov, A. R., Okuwaki, K., Otero-de-la-Roza, A., Parkin, S., Parunov, A., Podeszwa, R., Price, A. J. A., Price, L. S., Price, S. L., Probert, M. R., Pulido, A., Ramteke, G. R., Rehman, A. U., Reutzel-Edens, S. M., Rogal, J., Ross, M. J., Rumson, A. F., Sadiq, G., Saeed, Z. M., Salimi, A., Sasikumar, K., Sekharan, S., Shankland, K., Shi, B., Shi, X., Shinohara, K., Skillman, A. G., Song, H., Strasser, N., van de Streek, J., Sugden, I. J., Sun, G., Szalewicz, K., Tan, L., Tang, K., Tarczynski, F., Taylor, C. R., Tkatchenko, A., Tom, R., Touš, P., Tuckerman, M. E., Unzueta, P. A., Utsumi, Y., Vogt-Maranto, L., Weatherston, J., Wilkinson, L. J., Willacy, R. D., Wojtas, L., Woollam, G. R., Yang, Y., Yang, Z., Yonemochi, E., Yue, X., Zeng, Q., Zhou, T., Zhou, Y., Zubatyuk, R. & Cole, J. C. (2024). Acta Cryst. B80, 548-574.]). Overall, the DLPNO-CCSD(T1) conformational energy corrections are modest across the full set of crystal structures: the average magnitude of the shift is 1.1 kJ mol−1 relative to Form B, and the largest change is 2.9 kJ mol−1.

Next, we examine the role of the conformational energies in the relative lattice energies for the full set of candidate crystal structures more carefully. Because one of the crystal structures has Z′ = 2, the set of 99 crystal structures contains 100 symmetrically-unique monomer conformations. The gas-phase DLPNO-CCSD(T1) conformational energies of these structures span a roughly 15 kJ mol−1 range, with the conformations found in the GM and experimental structures mostly lying near the middle of the range (6.9–8.0 kJ mol−1 above the lowest-energy conformation). Amaj lies moderately higher at 11.1 kJ mol−1.

The key conformational flexibility in molecule XXXI involves the central dihedral angle d2 between the two rings, as defined in Fig. 1[link]. Secondary conformational flexibility involving dihedral angles d1 and d3 impacts the relative orientations of the two rings, with the fluorine on the benzene ring typically being either syn- or anti- relative to the isoxazoline heteroatoms. Fig. 4[link] superimposes the distribution of dihedral angles d2 for the set of monomer conformations onto gas-phase potential energy scans of d2 computed with B86bPBE-XDM and DLPNO-CCSD(T1). In 78% of the structures, molecule XXXI adopts an extended conformation with d2 lying in the range  ± 150–180°. This group includes forms Amaj, Amin, B, and the GM. Another 16% of the conformations adopt a more folded structure, with d2 dihedral angles of  ± 40–80°. These two clusters of dihedral angles generally lie near local energy minima of the gas-phase conformational energy profile. The remaining 6% of monomers have intermediate d2 angles in the  ± 120–140° range, which correspond to conformations that are energetically-unfavorable in the gas phase.

[Figure 4]
Figure 4
Gas-phase conformational energy profile for rotation about the central dihedral angle of molecule XXXI as computed with B86bPBE-XDM (purple) and DLPNO-CCSD(T1) (gray), and the difference between the two (green). For comparison, the figure also plots the energy differences between the DFT and coupled cluster models for each of the conformations found in the set of crystals as a function of the d2 dihedral angle (blue circles). For consistency with the crystal results in Fig. 2[link], all energies here are plotted relative to the molecular conformation found in the form B crystal (red circle).

None of these three molecule XXXI dihedral angles impacts the extent of π-conjugation in the molecule. Nevertheless, B86bPBE-XDM exhibits errors up to ∼5 kJ mol−1 relative to DLPNO-CCSD(T1) (Fig. 4[link]). The most notable discrepancies stem from B86bPBE-XDM underestimating the stability of the conformations with greater overlap of the ring systems. Similar behavior has been observed previously for B86bPBE-XDM and ππ interactions (Beran et al., 2022b[Beran, G. J. O., Wright, S. E., Greenwell, C. & Cruz-Cabeza, A. J. (2022b). J. Chem. Phys. 156, 104112.]). Although the crystal conformation energy differences between B86bPBE-XDM and DLPNO-CCSD(T1) are not completely captured by the dihedral angle d2 descriptor, Fig. 4[link] shows that B86bPBE-XDM errors in the crystal conformations (blue points) largely track with the difference between B86bPBE-XDM and DLPNO-CCSD(T1) computed on the gas-phase scan (green curve). The results for other electronic structure methods on this scan can be found in supporting information (Section S1.3).

Although B86bPBE-XDM predicts the XXXI conformational energies reasonably well, it is interesting to investigate what level of theory is required to achieve a more faithful description of the conformational energies for the moderately complex molecule XXXI that does not appear to exhibit strong π-conjugation delocalization error effects. Fig. 5[link] plots the error distributions for several different electronic structure models. Raw conformational energies are tabulated in supporting information (Section S1.2). In these box plots, the center line indicates the median error, the yellow box contains 50% of the errors, and the whiskers show the most extreme errors. Fig. 5[link] shows that both the B86bPBE-XDM and PBE-D4 GGA functionals perform similarly, with root-mean-square (rms) errors of 1.3 kJ mol−1. The hybrid functional PBE0-D4 reduces the rms error by 40% to 0.8 kJ mol−1. PBE0-D4 makes the largest improvements relative to the GGAs for the most strained conformations (d2 ∼ 120–140°) and near the d2∼ ± 60° basins. Unsurprisingly, the inclusion of van der Waals dispersion corrections is important for these conformational energies, and omitting the corrections increases the rms errors approximately two- to threefold (data not shown).

[Figure 5]
Figure 5
Gas-phase molecule XXXI conformational energy errors relative to DLPNO-CCSD(T1) for several electronic structure methods, in kJ mol−1. The root-mean-square error is also indicated. The conformational energies at each level of theory are defined relative to the average conformational energy in the set.

The range-separated hybrid meta-GGA ωB97M-V and the double-hybrid revDSD-PBEP86-D4 functionals perform even better than PBE0-D4, with rms errors of 0.4 kJ mol−1 and 0.3 kJ mol−1, respectively. MP2 performs well for the extended conformations which comprise most of the conformations in this set, but it overestimates the favorable ππ interactions between rings in the more-folded conformations, leading to a larger 0.7 kJ mol−1 rms error. The SCS-MP2D dispersion correction addresses this problem and reproduces the DLPNO-CCSD(T1) conformational energies with an rms error of only 0.2 kJ mol−1.

Examining the largest errors in the Fig. 5[link] conformational energy error distributions, the GGA errors lie within roughly  ± 3 kJ mol−1 of DLPNO-CCSD(T1), the global hybrid PBE0-D4 functional clearly improves the accuracy to around ± 2 kJ mol−1, the range-separated hybrid meta-GGA ωB97M-V and double-hybrid revDSD-PBEP86-D4 obtain errors within  ± 1 kJ mol−1, and SCS-MP2D achieves ± 0.6 kJ mol−1. The good fidelity among revDSD-PBEP86-D4, SCS-MP2D, and DLPNO-CCSD(T1) is further apparent in Fig. 6[link], which shows the strong correlations (R2 = 0.94–0.97) between the conformational energy corrections achieved when any of these three models is used to correct B86bPBE-XDM. As a result, a similar crystal energy landscape is obtained regardless of which of these methods (or even PBE0-D4) is used to correct the B86bPBE-XDM lattice energies [Fig. 2[link](a)]. The excellent consistency among the different methods also provides confidence in the accuracy of the benchmark conformational energies.

[Figure 6]
Figure 6
Correlations between the B86bPBE-XDM conformational energy corrections as computed with SCS-MP2D, revDSD-PBEP86-D4, and DLPNO-CCSD(T1) for molecules XXXI (red) and XXXII (blue). The conformational energy corrections were evaluated relative to form B of both species.

Finally, given the interest in using inexpensive semi-empirical models for intermediate refinement/ranking of structures, we tested three of Grimme's semi-empirical 3c methods: HF-3c, PBEh-3c, and r2SCAN-3c. HF-3c calculations are very fast but also rather unreliable, with an rms error of 2.8 kJ mol−1 and maximum errors up to 8 kJ mol−1 (supporting information Table S3). PBEh-3c and r2SCAN-3c offer better accuracy, with rms errors of 1.3 kJ mol−1 (max 4.0 kJ mol−1) and 1.1 kJ mol−1 (max 2.6 kJ mol−1), respectively (Fig. 5[link]). PBEh-3c is not recommended because it exhibits particularly large errors for certain conformations and is computationally expensive compared to the other 3c methods tested, but the excellent balance of low-computational cost and good accuracy of r2SCAN-3c is quite promising.

3.2. Molecule XXXII

Molecule XXXII is much larger than XXXI and has many more conformational degrees of freedom. It has a complex solid-form landscape, with at least eight anhydrate forms, four hydrates, six solvates, and seven other transient or unidentified forms having been observed. The blind test focused on the two anhydrates whose crystal structures have been determined, Forms A and B. The experimental screening report (Hunnisett et al., 2024[Hunnisett, L. M., Francia, N., Nyman, J., Abraham, N. S., Aitipamula, S., Alkhidir, T., Almehairbi, M., Anelli, A., Anstine, D. M., Anthony, J. E., Arnold, J. E., Bahrami, F., Bellucci, M. A., Beran, G. J. O., Bhardwaj, R. M., Bianco, R., Bis, J. A., Boese, A. D., Bramley, J., Braun, D. E., Butler, P. W. V., Cadden, J., Carino, S., Červinka, C., Chan, E. J., Chang, C., Clarke, S. M., Coles, S. J., Cook, C. J., Cooper, R. I., Darden, T., Day, G. M., Deng, W., Dietrich, H., DiPasquale, A., Dhokale, B., van Eijck, B. P., Elsegood, M. R. J., Firaha, D., Fu, W., Fukuzawa, K., Galanakis, N., Goto, H., Greenwell, C., Guo, R., Harter, J., Helfferich, J., Hoja, J., Hone, J., Hong, R., Hušák, M., Ikabata, Y., Isayev, O., Ishaque, O., Jain, V., Jin, Y., Jing, A., Johnson, E. R., Jones, I., Jose, K. V. J., Kabova, E. A., Keates, A., Kelly, P. F., Klimeš, J., Kostková, V., Li, H., Lin, X., List, A., Liu, C., Liu, Y. M., Liu, Z., Lončarić, I., Lubach, J. W., Ludík, J., Marom, N., Matsui, H., Mattei, A., Mayo, R. A., Melkumov, J. W., Mladineo, B., Mohamed, S., Momenzadeh Abardeh, Z., Muddana, H. S., Nakayama, N., Nayal, K. S., Neumann, M. A., Nikhar, R., Obata, S., O'Connor, D., Oganov, A. R., Okuwaki, K., Otero-de-la-Roza, A., Parkin, S., Parunov, A., Podeszwa, R., Price, A. J. A., Price, L. S., Price, S. L., Probert, M. R., Pulido, A., Ramteke, G. R., Rehman, A. U., Reutzel-Edens, S. M., Rogal, J., Ross, M. J., Rumson, A. F., Sadiq, G., Saeed, Z. M., Salimi, A., Sasikumar, K., Sekharan, S., Shankland, K., Shi, B., Shi, X., Shinohara, K., Skillman, A. G., Song, H., Strasser, N., van de Streek, J., Sugden, I. J., Sun, G., Szalewicz, K., Tan, L., Tang, K., Tarczynski, F., Taylor, C. R., Tkatchenko, A., Tom, R., Touš, P., Tuckerman, M. E., Unzueta, P. A., Utsumi, Y., Vogt-Maranto, L., Weatherston, J., Wilkinson, L. J., Willacy, R. D., Wojtas, L., Woollam, G. R., Yang, Y., Yang, Z., Yonemochi, E., Yue, X., Zeng, Q., Zhou, T., Zhou, Y., Zubatyuk, R. & Cole, J. C. (2024). Acta Cryst. B80, 548-574.]) provided to the blind test organizers indicates that Form B (#232, [P\bar{1}], Z = 4, Z′ = 2) is believed to be the thermo­dynamically stable polymorph, while Form A (#317, [P\bar{1}], Z = 2, Z′ = 1) is thermodynamically metastable, at least at room temperature and above. Form A exhibits disorder in rotation of the difluoro-methyl group, and structure #317 corresponds to the major component of the disorder, Amaj. It should be noted that the experimental investigation of the molecule XXXII solid-form landscape was challenging, and this understanding may be incomplete (Hunnisett et al., 2024[Hunnisett, L. M., Francia, N., Nyman, J., Abraham, N. S., Aitipamula, S., Alkhidir, T., Almehairbi, M., Anelli, A., Anstine, D. M., Anthony, J. E., Arnold, J. E., Bahrami, F., Bellucci, M. A., Beran, G. J. O., Bhardwaj, R. M., Bianco, R., Bis, J. A., Boese, A. D., Bramley, J., Braun, D. E., Butler, P. W. V., Cadden, J., Carino, S., Červinka, C., Chan, E. J., Chang, C., Clarke, S. M., Coles, S. J., Cook, C. J., Cooper, R. I., Darden, T., Day, G. M., Deng, W., Dietrich, H., DiPasquale, A., Dhokale, B., van Eijck, B. P., Elsegood, M. R. J., Firaha, D., Fu, W., Fukuzawa, K., Galanakis, N., Goto, H., Greenwell, C., Guo, R., Harter, J., Helfferich, J., Hoja, J., Hone, J., Hong, R., Hušák, M., Ikabata, Y., Isayev, O., Ishaque, O., Jain, V., Jin, Y., Jing, A., Johnson, E. R., Jones, I., Jose, K. V. J., Kabova, E. A., Keates, A., Kelly, P. F., Klimeš, J., Kostková, V., Li, H., Lin, X., List, A., Liu, C., Liu, Y. M., Liu, Z., Lončarić, I., Lubach, J. W., Ludík, J., Marom, N., Matsui, H., Mattei, A., Mayo, R. A., Melkumov, J. W., Mladineo, B., Mohamed, S., Momenzadeh Abardeh, Z., Muddana, H. S., Nakayama, N., Nayal, K. S., Neumann, M. A., Nikhar, R., Obata, S., O'Connor, D., Oganov, A. R., Okuwaki, K., Otero-de-la-Roza, A., Parkin, S., Parunov, A., Podeszwa, R., Price, A. J. A., Price, L. S., Price, S. L., Probert, M. R., Pulido, A., Ramteke, G. R., Rehman, A. U., Reutzel-Edens, S. M., Rogal, J., Ross, M. J., Rumson, A. F., Sadiq, G., Saeed, Z. M., Salimi, A., Sasikumar, K., Sekharan, S., Shankland, K., Shi, B., Shi, X., Shinohara, K., Skillman, A. G., Song, H., Strasser, N., van de Streek, J., Sugden, I. J., Sun, G., Szalewicz, K., Tan, L., Tang, K., Tarczynski, F., Taylor, C. R., Tkatchenko, A., Tom, R., Touš, P., Tuckerman, M. E., Unzueta, P. A., Utsumi, Y., Vogt-Maranto, L., Weatherston, J., Wilkinson, L. J., Willacy, R. D., Wojtas, L., Woollam, G. R., Yang, Y., Yang, Z., Yonemochi, E., Yue, X., Zeng, Q., Zhou, T., Zhou, Y., Zubatyuk, R. & Cole, J. C. (2024). Acta Cryst. B80, 548-574.]).

From the initial set of 500 molecule XXXII crystal structures, 53 low-energy structures were selected for full B86bPBE-XDM geometry refinement as described in Section 2.1[link]. The resulting crystal energy landscapes computed with B86bPBE-XDM and three different conformational energy correction models are plotted in Fig. 7[link], and selected crystal structures are shown in Fig. 8[link]. At the B86bPBE-XDM level, forms Amaj and B lie 0.1 kJ mol−1 apart and 4.7–4.8 kJ mol−1 above the lowest-energy structure (#423). Applying either SCS-MP2D or revDSD-PBEP86-D4 conformational energy corrections moderately alters the B86bPBE-XDM landscape, stabilizing some of the lowest-energy structures relative to Amaj and B by a few tenths of a kJ mol−1, and making structure #500 (GM) similar to SCS-MP2D or lower (revDSD-PBEP86-D4) in energy than #423. However, applying the DLPNO-CCSD(T1) conformational correction alters the landscape considerably. Although forms Amaj and B lie within 2 kJ mol−1 of one another across all four landscapes, the DLPNO-CCSD(T1) conformational correction stabilizes a number of other low-energy structures substantially. As a result, Amaj now lies 8.8 kJ mol−1 above the GM, while the experimentally-preferred form B lies 10.5 kJ mol−1 above it. In other words, the DLPNO-CCSD(T1) conformational energy correction has a substantial impact, approximately doubling the energy window that separates the experimental forms from the most stable form on the landscape.

[Figure 7]
Figure 7
(a) Impact of applying intramolecular conformational energy corrections to the relative B86bPBE-XDM lattice energies of the candidate crystal structures for molecule XXXII. (b) Final B86bPBE-XDM + ΔDLPNO-CCSD(T1) crystal energy landscape for molecule XXXII. Form B is the most stable form experimentally. Tabulated energetics and additional CSP landscapes can be found in supporting information Section S2.1.
[Figure 8]
Figure 8
Crystal structures of the molecule XXXII polymorphs Amaj and B, and the predicted GM structure. Green arrows point to key dihedral angle d2, which is 108° in the GM structure versus nearly planar in the experimental polymorphs.

Some of the lower-energy structures on the crystal landscape might conceivably correspond to some of the other six uncharacterized anhydrate forms that have been observed experimentally. Nevertheless, it is surprising that the reportedly experimentally-preferred form B would be computed to lie 10.5 kJ mol−1 above the GM. The vast majority of observed crystal polymorphs (including conformational polymorphs) lie within 10 kJ mol−1 of one another (Nyman & Day, 2015[Nyman, J. & Day, G. M. (2015). CrystEngComm, 17, 5154-5165.]; Cruz-Cabeza et al., 2015[Cruz-Cabeza, A. J., Reutzel-Edens, S. M. & Bernstein, J. (2015). Chem. Soc. Rev. 44, 8619-8635.]), and the errors in the relative lattice energies computed with quantum chemistry are typical considerably smaller than 10 kJ mol−1 (Whittleton et al., 2017a[Whittleton, S. R., Otero-de-la-Roza, A. & Johnson, E. R. (2017a). J. Chem. Theory Comput. 13, 441-450.],b[Whittleton, S. R., Otero-de-la-Roza, A. & Johnson, E. R. (2017b). J. Chem. Theory Comput. 13, 5332-5342.]; Hoja et al., 2019[Hoja, J., Ko, H., Neumann, M. A., Car, R., DiStasio, R. A. Jr & Tkatchenko, A. (2019). Sci. Adv. 5, eaau3338.]; Greenwell et al., 2020[Greenwell, C., McKinley, J. L., Zhang, P., Zeng, Q., Sun, G., Li, B., Wen, S. & Beran, G. J. O. (2020). Chem. Sci. 11, 2200-2214.]; Beran et al., 2022a[Beran, G. J. O., Sugden, I. J., Greenwell, C., Bowskill, D. H., Pantelides, C. C. & Adjiman, C. S. (2022a). Chem. Sci. 13, 1288-1297.]).

The large differences between the final DLPNO-CCSD(T1)-corrected molecule XXXII landscape and the experimental understanding suggest that there are likely problems with the computational and/or experimental results. Potential issues might include the neglect of finite-temperature contributions that significantly reorder the structures on the landscape, unusually large errors in the quantum chemistry lattice energy calculations, and/or the incomplete experimental understanding of the system. Addressing any gaps in the experimental understanding is beyond the scope of this work. We have not performed free energy calculations on the crystals due to the large unit-cell sizes. However, a couple other participating groups in the blind test did perform free energy calculations, and forms Amaj and B remained at least several kJ mol−1 above those groups' respective GM structures (Hunnisett et al., 2024[Hunnisett, L. M., Francia, N., Nyman, J., Abraham, N. S., Aitipamula, S., Alkhidir, T., Almehairbi, M., Anelli, A., Anstine, D. M., Anthony, J. E., Arnold, J. E., Bahrami, F., Bellucci, M. A., Beran, G. J. O., Bhardwaj, R. M., Bianco, R., Bis, J. A., Boese, A. D., Bramley, J., Braun, D. E., Butler, P. W. V., Cadden, J., Carino, S., Červinka, C., Chan, E. J., Chang, C., Clarke, S. M., Coles, S. J., Cook, C. J., Cooper, R. I., Darden, T., Day, G. M., Deng, W., Dietrich, H., DiPasquale, A., Dhokale, B., van Eijck, B. P., Elsegood, M. R. J., Firaha, D., Fu, W., Fukuzawa, K., Galanakis, N., Goto, H., Greenwell, C., Guo, R., Harter, J., Helfferich, J., Hoja, J., Hone, J., Hong, R., Hušák, M., Ikabata, Y., Isayev, O., Ishaque, O., Jain, V., Jin, Y., Jing, A., Johnson, E. R., Jones, I., Jose, K. V. J., Kabova, E. A., Keates, A., Kelly, P. F., Klimeš, J., Kostková, V., Li, H., Lin, X., List, A., Liu, C., Liu, Y. M., Liu, Z., Lončarić, I., Lubach, J. W., Ludík, J., Marom, N., Matsui, H., Mattei, A., Mayo, R. A., Melkumov, J. W., Mladineo, B., Mohamed, S., Momenzadeh Abardeh, Z., Muddana, H. S., Nakayama, N., Nayal, K. S., Neumann, M. A., Nikhar, R., Obata, S., O'Connor, D., Oganov, A. R., Okuwaki, K., Otero-de-la-Roza, A., Parkin, S., Parunov, A., Podeszwa, R., Price, A. J. A., Price, L. S., Price, S. L., Probert, M. R., Pulido, A., Ramteke, G. R., Rehman, A. U., Reutzel-Edens, S. M., Rogal, J., Ross, M. J., Rumson, A. F., Sadiq, G., Saeed, Z. M., Salimi, A., Sasikumar, K., Sekharan, S., Shankland, K., Shi, B., Shi, X., Shinohara, K., Skillman, A. G., Song, H., Strasser, N., van de Streek, J., Sugden, I. J., Sun, G., Szalewicz, K., Tan, L., Tang, K., Tarczynski, F., Taylor, C. R., Tkatchenko, A., Tom, R., Touš, P., Tuckerman, M. E., Unzueta, P. A., Utsumi, Y., Vogt-Maranto, L., Weatherston, J., Wilkinson, L. J., Willacy, R. D., Wojtas, L., Woollam, G. R., Yang, Y., Yang, Z., Yonemochi, E., Yue, X., Zeng, Q., Zhou, T., Zhou, Y., Zubatyuk, R. & Cole, J. C. (2024). Acta Cryst. B80, 548-574.]).

Here, we investigate one aspect of the accuracy of the lattice energy calculations: the accuracy of the conformational energies. Molecule XXXII exhibits a wide variety of conformations across the 55 symmetrically-unique molecular structures extracted from the 53 crystal structures. They range from highly-extended to somewhat folded, and span a gas-phase energy window of more than 40 kJ mol−1. The most stable conformations adopt highly-folded structures with strong intramolecular interactions. Interestingly, the GM crystal structure contains the least stable conformation in this set, lying 43 kJ mol−1 above the folded conformation in structure #331 in the gas phase. In contrast, the conformational energies of the experimental polymorphs are approximately average within the set (supporting information Section S2.2). In the condensed phase, extended conformations will be less penalized due to the ability to form intermolecular interactions with the surrounding environment. For example, placing the models in a methanol or water polarizable continuum solvent model reduces the energy window spanned by the conformations moderately, but the GM conformation still remains among the least stable conformations and lies ∼25 kJ mol−1 above the structure #331 conformation or ∼10–20 kJ mol−1 above the form Amaj or B conformations.

Among the many conformational degrees of freedom in molecule XXXII, dihedral angles d1d4 (Fig. 1[link]) have the largest potential impact on the extent of π-conjugation and therefore to manifest delocalization error issues. For d1, all 55 conformers examined adopt a narrow range of angles near 45° (or 135°). The absence of significant variations in the extent of π-conjugation across d1 among the polymorphs means that any errors in describing the energetics associated with dihedral d1 should largely cancel in the relative lattice energies. Similarly, the intramolecular N—H⋯N hydrogen bond prevents dihedral angle d3 from deviating significantly from planarity in this set of structures. However, dihedral angles d2 and d4 vary more widely across the different structures and prove to be the most important with regard to π-conjugation-related delocalization error in practice.

Specifically, whereas most of the molecule XXXII conformations adopt roughly planar d2 and d4 angles, eight structures rotate the amide torsion d2 out of the plane to ∼100–110°, and three rotate the thioether out of the plane (d4) to ∼40–65°. Both changes decrease the amount of π-conjugation in the molecule, making these structures potentially problematic for density functionals that exhibit substantial delocalization error. To test this, Fig. 9[link] plots conformational energy scans performed for d2 and d4 on the fragments of molecule XXXII shown. Similar scans for dihedral angles d1 and d3 are provided in the supporting information Section S2.3. For both d2 and d4, GGA functionals B86bPBE-XDM and PBE-D4 overestimate the torsional barriers by up to ∼4–5 kJ mol−1 compared to DLPNO-CCSD(T1) – i.e. they over-stabilize the more conjugated planar structures. Problematic delocalization error issues have been observed for both of these functional group types previously (Beran et al., 2022b[Beran, G. J. O., Wright, S. E., Greenwell, C. & Cruz-Cabeza, A. J. (2022b). J. Chem. Phys. 156, 104112.]). The hybrid PBE0-D4 functional reduces the error somewhat, while further improvements are obtained with the higher-level methods. Interestingly, whereas most of the models tested here overestimate the torsion barriers, ωB97M-V underestimates them. SCS-MP2D performs very well for dihedral d4, but it overestimates the barrier for dihedral d2 by up to 2 kJ mol−1, versus only 1 kJ mol−1 for revDSD-PBEP86-D4. Note that in contrast to d4, the orientation of the –OCHF2 group para- to the thioether exhibits much smaller B86bPBE-XDM conformational energy errors.

[Figure 9]
Figure 9
Potential energy scans for key dihedral angles d2 and d4 using the molecular fragments of molecule XXXII are shown. The upper panel shows the potential energy curve, while the lower one plots the errors relative to the DLPNO-CCSD(T1).

The errors observed for the 11 aforementioned molecule XXXII crystal conformations that deviate from planarity about d2 or d4 are consistent with these fragment molecule scans. B86bPBE-XDM and PBE-D4 overestimate the DLPNO-CCSD(T1) conformational energies for the conformations with non-planar d2 and d4 by an average 4.6 and 5.2 kJ mol−1, respectively, with maximum errors of nearly 7 kJ mol−1. Switching to the hybrid PBE0-D4 functional reduces that to 3.3 kJ mol−1 (max 4.6 kJ mol−1), SCS-MP2D to 2.5 kJ mol−1 (max 4.2 kJ mol−1) and revDSD-PBEP86-D4 to 2.2 kJ mol−1 (max 3.7 kJ mol−1). Returning to the crystal polymorph energies in Fig. 7[link], almost all of the large disagreements between the SCS-MP2D and DLPNO-CCSD(T)-corrected lattice energies occur for structures that exhibit non-planar amide bonds d2. DLPNO-CCSD(T1) stabilizes conformations with non-planar amides more so than SCS-MP2D or revDSD-PBEP86-D4. As a result, B86bPBE-XDM with DLPNO-CCSD(T1) conformational energy corrections predicts those non-planar amide structures to be the most stable ones on the entire landscape (Figs. 7[link] and 8[link]).

In addition to the conformations that are under-stabilized by B86bPBE-XDM, seven other molecular XXXII conformations are over-stabilized by ∼3–5 kJ mol−1 relative to DLPNO-CCSD(T1) (3.9 kJ mol−1 on average, see supporting information Table S6 for details). No unifying structural trends were identified among these structures. The largest-error case adopts a fairly folded conformation, while others are more extended. Errors associated with the conformational energies of the saturated six-membered rings also appear to play a role in some cases. These cases serve as a reminder that the conformational energies of highly flexible molecules can be difficult to model correctly even without changes in π-conjugation.

Fig. 10[link] plots the conformational energy error distributions obtained using various electronic structure models. Similar to molecule XXXI, improving the electronic structure model generally reduces the errors relative to DLPNO-CCSD(T1). The 1.4–1.7 kJ mol−1 rms errors for the three best models, ωB97M-V, SCS-MP2D, and revDSD-PBEP86-D4 are about half as large as those for the GGA functionals (2.8–3.1 kJ mol−1). Hybrid PBE0-D4 lies in between the two sets with an rms error of 2.3 kJ mol−1. Among the 3c methods, only r2SCAN-3c gives reasonable accuracy, with an rms error of 2.4 kJ mol−1, slightly larger than PBE0-D4. However, it is notable that the widths of the error distributions and the root-mean-square errors for molecule XXXII are substantially larger than for XXXI. For example, SCS-MP2D gives an rms error of only 0.2 kJ mol−1 for molecule XXXI, compared to 1.6 kJ mol−1 for XXXII. Moreover, the conformational energy corrections computed with SCS-MP2D, revDSD-PBEP86-D4, and DLPNO-CCSD(T1) all disagree considerably with one another (Fig. 6[link]), with R2 values of 0.56–0.72. This markedly contrasts the high consistency and excellent R2 values of 0.94–0.97 found for molecule XXXI.

[Figure 10]
Figure 10
Gas-phase molecule XXXII conformational energy errors (kJ mol−1) versus DLPNO-CCSD(T1) for several electronic structure methods. The conformational energies at each level of theory are defined relative to the average conformational energy in the set.

The large discrepancies between DLPNO-CCSD(T1) and most of the other methods might raise questions about the reliability of the DLPNO-CCSD(T1) benchmarks. As discussed in Section 2.3[link], however, test calculations suggest the DLPNO-CCSD(T1) relative conformational energies for most conformations appear to be converged to within ∼0.5 kJ mol−1 with regard to the DLPNO numerical thresholds, basis set, and use of T1 instead of T0 triples. The DLPNO-CCSD(T) calculations that proved most difficult to converge involved the highly folded conformations (especially #331 and #120), but the agreement among SCS-MP2D, revDSD-PBEP86-D4, and DLPNO-CCSD(T1) is reasonable for those conformations (supporting information Table S6 ). We further test the DLPNO-CCSD(T1) calculations by comparing them against canonical density-fitted CCSD(T)/aug-cc-pVDZ energy calculations for the small fragment molecule scans about d2 and d4. Those tests found that DLPNO-CCSD(T1) differs from full CCSD(T) by 0.4 kJ mol−1 or less for d2, and 0.1 kJ mol−1 or less for d4. This does not rule out the possibility that the local approximations in DLPNO-CCSD(T1) become more problematic in the full molecule, but the fact that the errors between SCS-MP2D and DLPNO-CCSD(T1) are similar in magnitude for both the small fragments and the full molecules suggests that the small fragments provide a reasonable model system. In other words, the DLPNO-CCSD(T1) energies appear robust across several different potential sources of error.

Due to the large size of molecule XXXII, we have not benchmarked the performance of the B86bPBE-XDM intermolecular interactions against higher-level electronic structure methods. However, we observe that the energy difference between forms A and B and the energies of those forms relative to the GM obtained with our conformationally-corrected B86bPBE-XDM results appear to be generally consistent with the results from other participating groups in the blind test that applied hybrid density functionals to the full crystals (Hunnisett et al., 2024[Hunnisett, L. M., Francia, N., Nyman, J., Abraham, N. S., Aitipamula, S., Alkhidir, T., Almehairbi, M., Anelli, A., Anstine, D. M., Anthony, J. E., Arnold, J. E., Bahrami, F., Bellucci, M. A., Beran, G. J. O., Bhardwaj, R. M., Bianco, R., Bis, J. A., Boese, A. D., Bramley, J., Braun, D. E., Butler, P. W. V., Cadden, J., Carino, S., Červinka, C., Chan, E. J., Chang, C., Clarke, S. M., Coles, S. J., Cook, C. J., Cooper, R. I., Darden, T., Day, G. M., Deng, W., Dietrich, H., DiPasquale, A., Dhokale, B., van Eijck, B. P., Elsegood, M. R. J., Firaha, D., Fu, W., Fukuzawa, K., Galanakis, N., Goto, H., Greenwell, C., Guo, R., Harter, J., Helfferich, J., Hoja, J., Hone, J., Hong, R., Hušák, M., Ikabata, Y., Isayev, O., Ishaque, O., Jain, V., Jin, Y., Jing, A., Johnson, E. R., Jones, I., Jose, K. V. J., Kabova, E. A., Keates, A., Kelly, P. F., Klimeš, J., Kostková, V., Li, H., Lin, X., List, A., Liu, C., Liu, Y. M., Liu, Z., Lončarić, I., Lubach, J. W., Ludík, J., Marom, N., Matsui, H., Mattei, A., Mayo, R. A., Melkumov, J. W., Mladineo, B., Mohamed, S., Momenzadeh Abardeh, Z., Muddana, H. S., Nakayama, N., Nayal, K. S., Neumann, M. A., Nikhar, R., Obata, S., O'Connor, D., Oganov, A. R., Okuwaki, K., Otero-de-la-Roza, A., Parkin, S., Parunov, A., Podeszwa, R., Price, A. J. A., Price, L. S., Price, S. L., Probert, M. R., Pulido, A., Ramteke, G. R., Rehman, A. U., Reutzel-Edens, S. M., Rogal, J., Ross, M. J., Rumson, A. F., Sadiq, G., Saeed, Z. M., Salimi, A., Sasikumar, K., Sekharan, S., Shankland, K., Shi, B., Shi, X., Shinohara, K., Skillman, A. G., Song, H., Strasser, N., van de Streek, J., Sugden, I. J., Sun, G., Szalewicz, K., Tan, L., Tang, K., Tarczynski, F., Taylor, C. R., Tkatchenko, A., Tom, R., Touš, P., Tuckerman, M. E., Unzueta, P. A., Utsumi, Y., Vogt-Maranto, L., Weatherston, J., Wilkinson, L. J., Willacy, R. D., Wojtas, L., Woollam, G. R., Yang, Y., Yang, Z., Yonemochi, E., Yue, X., Zeng, Q., Zhou, T., Zhou, Y., Zubatyuk, R. & Cole, J. C. (2024). Acta Cryst. B80, 548-574.]). This provides cause for some optimism that the intermolecular energies associated with molecule XXXII may be easier to model than the conformational energies, and that the crystal lattice energy landscapes here are reasonably well-converged with respect to the quantum chemistry treatment.

Overall, the impact of conformational corrections beyond GGA DFT are much more important for molecule XXXII than for molecule XXXI. Even with the conformationally-corrected lattice energies, however, the predicted polymorph stabilities for molecule XXXII do not agree with the reported experimental stabilities of forms A and B, and the experimental forms lie surprisingly high above the predicted GM on the crystal energy landscape. Computing accurate conformational energies for molecule XXXII proves challenging even with state-of-the-art electronic structure methods. Nevertheless, between our own results and those from other groups participating in the blind test, we did not identify any evidence that the errors in the relative lattice energies would be large enough to account for the apparent disagreement between the computed landscape and the reported experimental interpretation. Further experimental work to solve additional crystal forms and to determine the crystal structures of the unknown forms would be very useful to resolving the discrepancies between theory and experiment. In closing, we note the possibility that it may be difficult to crystallize the GM and the other most stable crystal structures identified here, since the need to adopt the highly-unstable non-planar amide conformations could hinder their crystallization kinetics (Bhardwaj et al., 2019[Bhardwaj, R. M., McMahon, J. A., Nyman, J., Price, L. S., Konar, S., Oswald, I. D. H., Pulham, C. R., Price, S. L. & Reutzel-Edens, S. M. (2019). J. Am. Chem. Soc. 141, 13887-13897.]; Abramov et al., 2020[Abramov, Y. A., Zhang, P., Zeng, Q., Yang, M., Liu, Y. & Sekharan, S. (2020). Cryst. Growth Des. 20, 1512-1525.]).

4. Conclusions

The blind tests provide an excellent opportunity to assess the quality of state-of-the-art crystal structure prediction techniques. The seventh blind test contained several systems whose complexity is comparable to that frequently faced in industrial applications. Focusing on molecules XXXI and XXXII, we investigated the role of conformational energies in the crystal energy landscapes and sought to understand how well widely-used electronic structure methods can capture them. Molecule XXXI proves relatively straightforward: while typical GGA functionals exhibit modest errors in the conformational energies, these errors can be ameliorated readily using hybrid DFT or more advanced electronic structure models. Among the semi-empirical methods, r2SCAN-3c proves the most promising, with accuracy intermediate between the GGA and hybrid functionals at a low computational cost.

Molecule XXXII proves much more difficult, due to its extremely high conformational flexibility, wide range of conformational energies, and some examples of delocalization error derived from changes in the π-conjugation. Even the best range-separated and double-hybrid density functionals or dispersion-corrected MP2 methods tested here exhibit errors up to several kJ mol−1 relative to the DLPNO-CCSD(T1) conformational energies. Regardless, the results demonstrate the importance of refining the conformational energies beyond the GGA or even hybrid DFT levels for challenging CSP cases. The present study also demonstrates how consideration of multiple levels of theory in the energy calculations can help assess the errors and uncertainties associated with the predicted crystal energy landscapes.

Interestingly, for both molecules XXXI and XXXII, the relative lattice energies computed here disagree with what has been inferred from experiment. For XXXI, the discrepancies can probably be attributed to our neglect of finite temperature effects and/or the typical uncertainties associated with quantum chemistry calculations. For molecule XXXII, however, the discrepancies between the computed crystal energy landscape and the experimental understanding appear too large to be explained by such factors. Additional work to quantify the modeling errors associated with the intermolecular contribution to the lattice energies would be valuable. At the same time, further experimental efforts to solve the unknown molecule XXXII crystal structures and reveal their relative stabilities are clearly needed. Finally, further investigations should also consider whether the crystallization of the thermodynamically most-stable forms on the computational crystal energy landscape might be hindered by their highly-unfavorable intramolecular conformations.

Supporting information: crystal energy landscapes in graphical and tabulated forms, tables of the conformational energies, additional conformational energy scans, and data from the DLPNO-CCSD(T) convergence testing are provided in PDF format. B86bPBE-XDM optimized crystal structures and the molecular geometries extracted from the crystals are provided in CIF and XYZ formats, respectively.

Supporting information


Acknowledgements

All opinions expressed in this paper are the author's and do not necessarily reflect the policies and views of DOE, ORAU, or ORISE.

Funding information

GJOB gratefully acknowledges funding from the National Science Foundation (CHE-1955554) and supercomputer time from ACCESS (CHE110064). Computations were also performed using the computer clusters and data storage resources of the UC Riverside HPCC, which were funded by grants from NSF (MRI-2215705, MRI-1429826) and NIH (1S10OD016290-01A1). Additional support for this work to PAU came from the US Department of Energy, Office of Science, Office of Workforce Development for Teachers and Scientists, Office of Science Graduate Student Research (SCGSR) program. The SCGSR program is administered by the Oak Ridge Institute for Science and Education (ORISE) for the DOE. ORISE is managed by ORAU under contract number DE-SC0014664.

References

First citationAbramov, Y. A., Zhang, P., Zeng, Q., Yang, M., Liu, Y. & Sekharan, S. (2020). Cryst. Growth Des. 20, 1512–1525.  Web of Science CrossRef CAS Google Scholar
First citationAdamo, C. & Barone, V. (1999). J. Chem. Phys. 110, 6158–6170.  Web of Science CrossRef CAS Google Scholar
First citationBardwell, D. A., Adjiman, C. S., Arnautova, Y. A., Bartashevich, E., Boerrigter, S. X. M., Braun, D. E., Cruz-Cabeza, A. J., Day, G. M., Della Valle, R. G., Desiraju, G. R., van Eijck, B. P., Facelli, J. C., Ferraro, M. B., Grillo, D., Habgood, M., Hofmann, D. W. M., Hofmann, F., Jose, K. V. J., Karamertzanis, P. G., Kazantsev, A. V., Kendrick, J., Kuleshova, L. N., Leusen, F. J. J., Maleev, A. V., Misquitta, A. J., Mohamed, S., Needs, R. J., Neumann, M. A., Nikylov, D., Orendt, A. M., Pal, R., Pantelides, C. C., Pickard, C. J., Price, L. S., Price, S. L., Scheraga, H. A., van de Streek, J., Thakur, T. S., Tiwari, S., Venuti, E. & Zhitkov, I. K. (2011). Acta Cryst. B67, 535–551.  Web of Science CrossRef IUCr Journals Google Scholar
First citationBecke, A. D. (1986). J. Chem. Phys. 85, 7184–7187.  CrossRef CAS Web of Science Google Scholar
First citationBeran, G. J. O. (2019). CrystEngComm, 21, 758–764.  Web of Science CrossRef CAS Google Scholar
First citationBeran, G. J. O. (2023). Chem. Sci. 14, 13290–13312.  Web of Science CrossRef CAS PubMed Google Scholar
First citationBeran, G. J. O., Greenwell, C., Cook, C. & Řezáč, J. (2023). Acc. Chem. Res. 56, 3525–3534.  Web of Science CrossRef CAS PubMed Google Scholar
First citationBeran, G. J. O., Sugden, I. J., Greenwell, C., Bowskill, D. H., Pantelides, C. C. & Adjiman, C. S. (2022a). Chem. Sci. 13, 1288–1297.  Web of Science CrossRef CAS PubMed Google Scholar
First citationBeran, G. J. O., Wright, S. E., Greenwell, C. & Cruz-Cabeza, A. J. (2022b). J. Chem. Phys. 156, 104112.  Web of Science CrossRef PubMed Google Scholar
First citationBhardwaj, R. M., McMahon, J. A., Nyman, J., Price, L. S., Konar, S., Oswald, I. D. H., Pulham, C. R., Price, S. L. & Reutzel-Edens, S. M. (2019). J. Am. Chem. Soc. 141, 13887–13897.  Web of Science CSD CrossRef CAS PubMed Google Scholar
First citationBowskill, D. H., Sugden, I. J., Konstantinopoulos, S., Adjiman, C. S. & Pantelides, C. C. (2021). Annu. Rev. Chem. Biomol. Eng. 12, 593–623.  Web of Science CrossRef CAS PubMed Google Scholar
First citationBrandenburg, J. G. & Grimme, S. (2014). In Prediction and Calculation of Crystal Structures: Methods and Applications, Vol. 345, Topics in Current Chemistry, edited by S. Atahan-Evrenk & A. Aspuru-Guzik, pp. 1–24. Cham: Springer International Publishing.  Google Scholar
First citationBraun, D. E., Lingireddy, S. R., Beidelschies, M. D., Guo, R., Müller, P., Price, S. L. & Reutzel-Edens, S. M. (2017). Cryst. Growth Des. 17, 5349–5365.  Web of Science CSD CrossRef CAS PubMed Google Scholar
First citationBraun, D. E., McMahon, J. A., Bhardwaj, R. M., Nyman, J., Neumann, M. A., van de Streek, J. & Reutzel-Edens, S. M. (2019). Cryst. Growth Des. 19, 2947–2962.  Web of Science CSD CrossRef CAS Google Scholar
First citationCaldeweyher, E., Bannwarth, C. & Grimme, S. (2017). J. Chem. Phys. 147, 034112.  Web of Science CrossRef PubMed Google Scholar
First citationCook, C. J., Li, W., Lui, B. F., Gately, T. J., Al-Kaysi, R. O., Mueller, L. J., Bardeen, C. J. & Beran, G. J. O. (2023). Chem. Sci. 14, 937–949.  Web of Science CrossRef CAS PubMed Google Scholar
First citationCook, C. J., Perry, C. J. & Beran, G. J. O. (2023). J. Phys. Chem. Lett. 14, 6823–6831.  Web of Science CrossRef CAS PubMed Google Scholar
First citationCruz-Cabeza, A. J., Reutzel-Edens, S. M. & Bernstein, J. (2015). Chem. Soc. Rev. 44, 8619–8635.  Web of Science CAS PubMed Google Scholar
First citationCutini, M., Civalleri, B., Corno, M., Orlando, R., Brandenburg, J. G., Maschio, L. & Ugliengo, P. (2016). J. Chem. Theory Comput. 12, 3340–3352.  Web of Science CrossRef CAS PubMed Google Scholar
First citationDay, G. M., Cooper, T. G., Cruz-Cabeza, A. J., Hejczyk, K. E., Ammon, H. L., Boerrigter, S. X. M., Tan, J. S., Della Valle, R. G., Venuti, E., Jose, J., Gadre, S. R., Desiraju, G. R., Thakur, T. S., van Eijck, B. P., Facelli, J. C., Bazterra, V. E., Ferraro, M. B., Hofmann, D. W. M., Neumann, M. A., Leusen, F. J. J., Kendrick, J., Price, S. L., Misquitta, A. J., Karamertzanis, P. G., Welch, G. W. A., Scheraga, H. A., Arnautova, Y. A., Schmidt, M. U., van de Streek, J., Wolf, A. K. & Schweizer, B. (2009). Acta Cryst. B65, 107–125.  Web of Science CSD CrossRef IUCr Journals Google Scholar
First citationDay, G. M., Motherwell, W. D. S., Ammon, H. L., Boerrigter, S. X. M., Della Valle, R. G., Venuti, E., Dzyabchenko, A., Dunitz, J. D., Schweizer, B., van Eijck, B. P., Erk, P., Facelli, J. C., Bazterra, V. E., Ferraro, M. B., Hofmann, D. W. M., Leusen, F. J. J., Liang, C., Pantelides, C. C., Karamertzanis, P. G., Price, S. L., Lewis, T. C., Nowell, H., Torrisi, A., Scheraga, H. A., Arnautova, Y. A., Schmidt, M. U. & Verwer, P. (2005). Acta Cryst. B61, 511–527.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationDovesi, R., Erba, A., Orlando, R., Zicovich–Wilson, C. M., Civalleri, B., Maschio, L., Rérat, M., Casassa, S., Baima, J., Salustro, S. & Kirtman, B. (2018). WIREs Comput. Mol. Sci. 8, e1360.  Google Scholar
First citationDunning, T. H. (1989). J. Chem. Phys. 90, 1007–1023.  CrossRef CAS Web of Science Google Scholar
First citationFiraha, D., Liu, Y. M., van de Streek, J., Sasikumar, K., Dietrich, H., Helfferich, J., Aerts, L., Braun, D. E., Broo, A., DiPasquale, A. G., Lee, A. Y., Le Meur, S., Nilsson Lill, S. O., Lunsmann, W. J., Mattei, A., Muglia, P., Putra, O. D., Raoui, M., Reutzel-Edens, S. M., Rome, S., Sheikh, A. Y., Tkatchenko, A., Woollam, G. R. & Neumann, M. A. (2023). Nature, 623, 324–328.  Web of Science CrossRef CAS PubMed Google Scholar
First citationGately, T. J., Cook, C. C., Almuzarie, R., Islam, I., Gardner, Z. T., Iuliucci, R. J., Al-Kaysi, R. O., Beran, G. J. O. & Bardeen, C. J. (2022). Cryst. Growth Des. 22, 7298–7307.  Web of Science CSD CrossRef CAS Google Scholar
First citationGately, T. J., Sontising, W., Easley, C. J., Islam, I., Al-Kaysi, R. O., Beran, G. J. O. & Bardeen, C. J. (2021). CrystEngComm, 23, 5931–5943.  Web of Science CSD CrossRef CAS Google Scholar
First citationGiannozzi, P., Andreussi, O., Brumme, T., Bunau, O., Buongiorno Nardelli, M., Calandra, M., Car, R., Cavazzoni, C., Ceresoli, D., Cococcioni, M., Colonna, N., Carnimeo, I., Dal Corso, A., de Gironcoli, S., Delugas, P., DiStasio, R. A., Ferretti, A., Floris, A., Fratesi, G., Fugallo, G., Gebauer, R., Gerstmann, U., Giustino, F., Gorni, T., Jia, J., Kawamura, M., Ko, H.-Y., Kokalj, A., Küçükbenli, E., Lazzeri, M., Marsili, M., Marzari, N., Mauri, F., Nguyen, N. L., Nguyen, H.-V., Otero-de-la-Roza, A., Paulatto, L., Poncé, S., Rocca, D., Sabatini, R., Santra, B., Schlipf, M., Seitsonen, A. P., Smogunov, A., Timrov, I., Thonhauser, T., Umari, P., Vast, N., Wu, X. & Baroni, S. (2017). J. Phys. Condens. Matter, 29, 465901.  Web of Science CrossRef PubMed Google Scholar
First citationGreenwell, C. & Beran, G. J. O. (2018). The MP2D software library. https://github.com/Chandemonium/MP2D.  Google Scholar
First citationGreenwell, C. & Beran, G. J. O. (2020). Cryst. Growth Des. 20, 4875–4881.  Web of Science CrossRef CAS Google Scholar
First citationGreenwell, C. & Beran, G. J. O. (2021). J. Mater. Chem. C, 9, 2848–2857.  Web of Science CrossRef CAS Google Scholar
First citationGreenwell, C., McKinley, J. L., Zhang, P., Zeng, Q., Sun, G., Li, B., Wen, S. & Beran, G. J. O. (2020). Chem. Sci. 11, 2200–2214.  Web of Science CrossRef CAS PubMed Google Scholar
First citationGreenwell, C., Řezáč, J. & Beran, G. J. O. (2022). Phys. Chem. Chem. Phys. 24, 3695–3712.  Web of Science CrossRef CAS PubMed Google Scholar
First citationGrimme, S., Brandenburg, J. G., Bannwarth, C. & Hansen, A. (2015). J. Chem. Phys. 143, 054107.  Web of Science CrossRef PubMed Google Scholar
First citationGrimme, S., Hansen, A., Ehlert, S. & Mewes, J.-M. (2021). J. Chem. Phys. 154, 064103.  Web of Science CrossRef PubMed Google Scholar
First citationGuo, Y., Riplinger, C., Becker, U., Liakos, D. G., Minenkov, Y., Cavallo, L. & Neese, F. (2018). J. Chem. Phys. 148, 011101.  Web of Science CrossRef PubMed Google Scholar
First citationHelgaker, T., Klopper, W., Koch, H. & Noga, J. (1997). J. Chem. Phys. 106, 9639–9646.  CrossRef CAS Web of Science Google Scholar
First citationHoja, J., Ko, H., Neumann, M. A., Car, R., DiStasio, R. A. Jr & Tkatchenko, A. (2019). Sci. Adv. 5, eaau3338.  Web of Science CrossRef PubMed Google Scholar
First citationHunnisett, L. M., Francia, N., Nyman, J., Abraham, N. S., Aitipamula, S., Alkhidir, T., Almehairbi, M., Anelli, A., Anstine, D. M., Anthony, J. E., Arnold, J. E., Bahrami, F., Bellucci, M. A., Beran, G. J. O., Bhardwaj, R. M., Bianco, R., Bis, J. A., Boese, A. D., Bramley, J., Braun, D. E., Butler, P. W. V., Cadden, J., Carino, S., Červinka, C., Chan, E. J., Chang, C., Clarke, S. M., Coles, S. J., Cook, C. J., Cooper, R. I., Darden, T., Day, G. M., Deng, W., Dietrich, H., DiPasquale, A., Dhokale, B., van Eijck, B. P., Elsegood, M. R. J., Firaha, D., Fu, W., Fukuzawa, K., Galanakis, N., Goto, H., Greenwell, C., Guo, R., Harter, J., Helfferich, J., Hoja, J., Hone, J., Hong, R., Hušák, M., Ikabata, Y., Isayev, O., Ishaque, O., Jain, V., Jin, Y., Jing, A., Johnson, E. R., Jones, I., Jose, K. V. J., Kabova, E. A., Keates, A., Kelly, P. F., Klimeš, J., Kostková, V., Li, H., Lin, X., List, A., Liu, C., Liu, Y. M., Liu, Z., Lončarić, I., Lubach, J. W., Ludík, J., Marom, N., Matsui, H., Mattei, A., Mayo, R. A., Melkumov, J. W., Mladineo, B., Mohamed, S., Momenzadeh Abardeh, Z., Muddana, H. S., Nakayama, N., Nayal, K. S., Neumann, M. A., Nikhar, R., Obata, S., O'Connor, D., Oganov, A. R., Okuwaki, K., Otero-de-la-Roza, A., Parkin, S., Parunov, A., Podeszwa, R., Price, A. J. A., Price, L. S., Price, S. L., Probert, M. R., Pulido, A., Ramteke, G. R., Rehman, A. U., Reutzel-Edens, S. M., Rogal, J., Ross, M. J., Rumson, A. F., Sadiq, G., Saeed, Z. M., Salimi, A., Sasikumar, K., Sekharan, S., Shankland, K., Shi, B., Shi, X., Shinohara, K., Skillman, A. G., Song, H., Strasser, N., van de Streek, J., Sugden, I. J., Sun, G., Szalewicz, K., Tan, L., Tang, K., Tarczynski, F., Taylor, C. R., Tkatchenko, A., Tom, R., Touš, P., Tuckerman, M. E., Unzueta, P. A., Utsumi, Y., Vogt-Maranto, L., Weatherston, J., Wilkinson, L. J., Willacy, R. D., Wojtas, L., Woollam, G. R., Yang, Y., Yang, Z., Yonemochi, E., Yue, X., Zeng, Q., Zhou, T., Zhou, Y., Zubatyuk, R. & Cole, J. C. (2024). Acta Cryst. B80, 548–574.  Google Scholar
First citationKaramertzanis, P. G., Day, G. M., Welch, G. W. A., Kendrick, J., Leusen, F. J. J., Neumann, M. A. & Price, S. L. (2008). J. Chem. Phys. 128, 244708.  Web of Science CrossRef PubMed Google Scholar
First citationLeBlanc, L. M., Dale, S. G., Taylor, C. R., Becke, A. D., Day, G. M. & Johnson, E. R. (2018). Angew. Chem. Int. Ed. 57, 14906–14910.  Web of Science CrossRef CAS Google Scholar
First citationLommerse, J. P. M., Motherwell, W. D. S., Ammon, H. L., Dunitz, J. D., Gavezzotti, A., Hofmann, D. W. M., Leusen, F. J. J., Mooij, W. T. M., Price, S. L., Schweizer, B., Schmidt, M. U., van Eijck, B. P., Verwer, P. & Williams, D. E. (2000). Acta Cryst. B56, 697–714.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationMardirossian, N. & Head-Gordon, M. (2016). J. Chem. Phys. 144, 214110.  Web of Science CrossRef PubMed Google Scholar
First citationMartin, J. M. L. & Santra, G. (2020). Isr. J. Chem. 60, 787–804.  Web of Science CrossRef CAS Google Scholar
First citationMortazavi, M., Hoja, J., Aerts, L., Quéré, L., van de Streek, J., Neumann, M. A. & Tkatchenko, A. (2019). Commun. Chem. 2, 70.  Web of Science CrossRef Google Scholar
First citationMotherwell, W. D. S., Ammon, H. L., Dunitz, J. D., Dzyabchenko, A., Erk, P., Gavezzotti, A., Hofmann, D. W. M., Leusen, F. J. J., Lommerse, J. P. M., Mooij, W. T. M., Price, S. L., Scheraga, H., Schweizer, B., Schmidt, M. U., van Eijck, B. P., Verwer, P. & Williams, D. E. (2002). Acta Cryst. B58, 647–661.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationNeese, F. (2012). WIREs Comput. Mol. Sci. 2, 73–78.  Web of Science CrossRef CAS Google Scholar
First citationNeumann, M. A., Leusen, F. J. J. & Kendrick, J. (2008). Angew. Chem. Int. Ed. 47, 2427–2430.  Web of Science CrossRef CAS Google Scholar
First citationNeumann, M. A., van de Streek, J., Fabbiani, F. P. A., Hidber, P. & Grassmann, O. (2015). Nat. Commun. 6, 7793.   Google Scholar
First citationNikhar, R. & Szalewicz, K. (2022). Nat. Commun. 13, 3095.  Web of Science CrossRef PubMed Google Scholar
First citationNyman, J. & Day, G. M. (2015). CrystEngComm, 17, 5154–5165.  Web of Science CrossRef CAS Google Scholar
First citationNyman, J., Yu, L. & Reutzel-Edens, S. M. (2019). CrystEngComm, 21, 2080–2088.  Web of Science CSD CrossRef CAS Google Scholar
First citationOganov, A. R. (2018). Faraday Discuss. 211, 643–660.  Web of Science CrossRef CAS PubMed Google Scholar
First citationOtero-de-la-Roza, A. & Johnson, E. R. (2012). J. Chem. Phys. 136, 174109.  Web of Science PubMed Google Scholar
First citationOtero-de-la-Roza, A., LeBlanc, L. M. & Johnson, E. R. (2019). J. Chem. Theory Comput. 15, 4933–4944.  Web of Science CAS PubMed Google Scholar
First citationPerdew, J. P., Burke, K. & Ernzerhof, M. (1996). Phys. Rev. Lett. 77, 3865–3868.  CrossRef PubMed CAS Web of Science Google Scholar
First citationPerry, C. J. & Beran, G. J. O. (2023). Cryst. Growth Des. 23, 8352–8360.  Web of Science CrossRef CAS Google Scholar
First citationPrice, A. J., Mayo, R. A., Otero-de-la-Roza, A. & Johnson, E. R. (2023). CrystEngComm, 25, 953–960.  Web of Science CrossRef CAS Google Scholar
First citationPrice, A. J., Otero-de-la-Roza, A. & Johnson, E. R. (2022). Chem. Sci. 14, 1252–1262.  Web of Science CrossRef PubMed Google Scholar
First citationPrice, S. L. (2008). Int. Rev. Phys. Chem. 27, 541–568.  Web of Science CrossRef CAS Google Scholar
First citationReilly, A. M., Cooper, R. I., Adjiman, C. S., Bhattacharya, S., Boese, A. D., Brandenburg, J. G., Bygrave, P. J., Bylsma, R., Campbell, J. E., Car, R., Case, D. H., Chadha, R., Cole, J. C., Cosburn, K., Cuppen, H. M., Curtis, F., Day, G. M., DiStasio, R. A. Jr, Dzyabchenko, A., van Eijck, B. P., Elking, D. M., van den Ende, J. A., Facelli, J. C., Ferraro, M. B., Fusti-Molnar, L., Gatsiou, C.-A., Gee, T. S., de Gelder, R., Ghiringhelli, L. M., Goto, H., Grimme, S., Guo, R., Hofmann, D. W. M., Hoja, J., Hylton, R. K., Iuzzolino, L., Jankiewicz, W., de Jong, D. T., Kendrick, J., de Klerk, N. J. J., Ko, H.-Y., Kuleshova, L. N., Li, X., Lohani, S., Leusen, F. J. J., Lund, A. M., Lv, J., Ma, Y., Marom, N., Masunov, A. E., McCabe, P., McMahon, D. P., Meekes, H., Metz, M. P., Misquitta, A. J., Mohamed, S., Monserrat, B., Needs, R. J., Neumann, M. A., Nyman, J., Obata, S., Oberhofer, H., Oganov, A. R., Orendt, A. M., Pagola, G. I., Pantelides, C. C., Pickard, C. J., Podeszwa, R., Price, L. S., Price, S. L., Pulido, A., Read, M. G., Reuter, K., Schneider, E., Schober, C., Shields, G. P., Singh, P., Sugden, I. J., Szalewicz, K., Taylor, C. R., Tkatchenko, A., Tuckerman, M. E., Vacarro, F., Vasileiadis, M., Vazquez-Mayagoitia, A., Vogt, L., Wang, Y., Watson, R. E., de Wijs, G. A., Yang, J., Zhu, Q. & Groom, C. R. (2016). Acta Cryst. B72, 439–459.  Web of Science CrossRef IUCr Journals Google Scholar
First citationŘezáč, J. (2020a). J. Chem. Theory Comput. 16, 2355–2368.  Web of Science PubMed Google Scholar
First citationŘezáč, J. (2020b). J. Chem. Theory Comput. 16, 6305–6316.  Web of Science PubMed Google Scholar
First citationŘezáč, J. (2022). Phys. Chem. Chem. Phys. 24, 14780–14793.  Web of Science PubMed Google Scholar
First citationRiplinger, C., Pinski, P., Becker, U., Valeev, E. F. & Neese, F. (2016). J. Chem. Phys. 144, 024109.  Web of Science CrossRef PubMed Google Scholar
First citationRiplinger, C., Sandhoefer, B., Hansen, A. & Neese, F. (2013). J. Chem. Phys. 139, 134101.  Web of Science CrossRef PubMed Google Scholar
First citationSantra, G., Sylvetsky, N. & Martin, J. M. L. (2019). J. Phys. Chem. A, 123, 5129–5143.  Web of Science CrossRef CAS PubMed Google Scholar
First citationSmith, D. G. A., Burns, L. A., Simmonett, A. C., Parrish, R. M., Schieber, M. C., Galvelis, R., Kraus, P., Kruse, H., Di Remigio, R., Alenaizan, A., James, A. M., Lehtola, S., Misiewicz, J. P., Scheurer, M., Shaw, R. A., Schriber, J. B., Xie, Y., Glick, Z. L., Sirianni, D. A., O'Brien, J. S., Waldrop, J. M., Kumar, A., Hohenstein, E. G., Pritchard, B. P., Brooks, B. R., Schaefer, H. F., Sokolov, A. Y., Patkowski, K., DePrince, A. E., Bozkaya, U., King, R. A., Evangelista, F. A., Turney, J. M., Crawford, T. D. & Sherrill, C. D. (2020). J. Chem. Phys. 152, 184108.  Web of Science CrossRef PubMed Google Scholar
First citationSure, R. & Grimme, S. (2013). J. Comput. Chem. 34, 1672–1685.  Web of Science CrossRef CAS PubMed Google Scholar
First citationTan, M., Shtukenberg, A. G., Zhu, S., Xu, W., Dooryhee, E., Nichols, S. M., Ward, M. D., Kahr, B. & Zhu, Q. (2018). Faraday Discuss. 211, 477–491.  Web of Science CSD CrossRef CAS PubMed Google Scholar
First citationThomas, S. P. & Spackman, M. A. (2018). Aust. J. Chem. 71, 279.  CrossRef Google Scholar
First citationThompson, H. P. G. & Day, G. M. (2014). Chem. Sci. 5, 3173–3182.  Web of Science CrossRef CAS Google Scholar
First citationWhittleton, S. R., Otero-de-la-Roza, A. & Johnson, E. R. (2017a). J. Chem. Theory Comput. 13, 441–450.  Web of Science CrossRef CAS PubMed Google Scholar
First citationWhittleton, S. R., Otero-de-la-Roza, A. & Johnson, E. R. (2017b). J. Chem. Theory Comput. 13, 5332–5342.  Web of Science CrossRef CAS PubMed Google Scholar

This article is published by the International Union of Crystallography. Prior permission is not required to reproduce short quotations, tables and figures from this article, provided the original authors and source are cited. For more information, click here.

Journal logoSTRUCTURAL SCIENCE
CRYSTAL ENGINEERING
MATERIALS
ISSN: 2052-5206
Follow Acta Cryst. B
Sign up for e-alerts
Follow Acta Cryst. on Twitter
Follow us on facebook
Sign up for RSS feeds