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

ISSN: 2052-2525

Solid-state calculations for iterative refinement in quantum crystallography using the multipole model

crossmark logo

aChemische Kristallographie und Elektronenmikroskopie, Max-Planck-Institut für Kohlenforschung, Kaiser-Wilhelm-Platz 1, Mülheim an der Ruhr, 45470 North Rhine-Westphalia, Germany
*Correspondence e-mail: patzer@kofo.mpg.de

Edited by A. Fitch, ESRF, France (Received 15 November 2024; accepted 4 March 2025; online 4 April 2025)

This article is part of a collection of articles on Quantum Crystallography, and commemorates the 100th anniversary of the development of Quantum Mechanics.

A quantum crystallographic refinement methodology has been developed using theoretical multipole parameters generated directly from solid-state calculations using the CRYSTAL17 program. This refinement method is comparable to other transferable form factor approaches, such as the Invariom model, but in contrast to the Hirshfeld atom refinement, it uses theoretical multipole parameters to describe the electron density from a solid-state calculation performed with CRYSTAL17 in an iterative refinement procedure. For this purpose, a Python3 code named ReCrystal has been developed. To start ReCrystal, a CIF, a Gaussian basis set, a DFT functional and the number of CPUs must be defined. The Pack–Monkhorst and Gilat shrinking factors, which define a lattice in the first Brillouin zone, must also be specified. After k-point sampling, CRYSTAL17 calculates structure factors directly from the static electron density. Multipole parameters are generated from these structure factors using the XD program and are fixed during least-squares refinement. The refinement of the xylitol molecular crystal has shown that the hydrogen atom positions can be determined with reasonable agreement to those obtained in the neutron diffraction experiment. This indicates that the periodic boundary condition in ReCrystal is an improvement over gas phase refinement with HAR. The multipole parameters obtained from ReCrystal can be used for further charge density studies especially if weak interactions are the focus. In this work, we demonstrate the performance of ReCrystal on molecular crystals of the small molecules D/L-serine and xylitol with weak hydrogen-bonding motifs using multipole refinement. The advantage of this approach is that multipole parameters can be obtained from high-resolution calculated diffraction data, no database is required, and errors due to the model and errors resulting from the experiment are clearly separated.

1. Introduction

Quantum crystallographic (QCr) refinement techniques for single-crystal X-ray diffraction have undergone a revolutionary development in the last decades and have made a fundamental contribution to the understanding of the diffraction experiment in single-crystal structure analysis (Grabowsky et al., 2017[Grabowsky, S., Genoni, A. & Bürgi, H.-B. (2017). Chem. Sci. 8, 4159-4176.]; Genoni et al., 2018[Genoni, A., Bučinský, L., Claiser, N., Contreras-García, J., Dittrich, B., Dominiak, P. M., Espinosa, E., Gatti, C., Giannozzi, P., Gillet, J.-M., Jayatilaka, D., Macchi, P., Madsen, A. Ø., Massa, L., Matta, C. F., Merz, K. M. Jr, Nakashima, P. N. H., Ott, H., Ryde, U., Schwarz, K., Sierka, M. & Grabowsky, S. (2018). Chem. A Eur. J. 24, 10881-10905.]). These methods have led to a remarkable improvement of the refinement results due to the use of aspherical atomic form factors (Dittrich et al., 2005[Dittrich, B., Hübschle, C. B., Messerschmidt, M., Kalinowski, R., Girnt, D. & Luger, P. (2005). Acta Cryst. A61, 314-320.]; Dittrich et al., 2006[Dittrich, B., Strumpel, M., Schäfer, M., Spackman, M. A. & Koritsánszky, T. (2006). Acta Cryst. A62, 217-223.]; Dominiak et al., 2007[Dominiak, P. M., Volkov, A., Li, X., Messerschmidt, M. & Coppens, P. (2007). J. Chem. Theory Comput. 3, 232-247.]; Jayatilaka & Dittrich, 2008[Jayatilaka, D. & Dittrich, B. (2008). Acta Cryst. A64, 383-393.]; Capelli et al., 2014[Capelli, S. C., Bürgi, H.-B., Dittrich, B., Grabowsky, S. & Jayatilaka, D. (2014). IUCrJ, 1, 361-379.]). Single-crystal structure analysis applies a model of the charge density distribution in the unit cell to obtain information about the phases of the intensities in the diffraction pattern. This model is related to the diffraction intensities via a Fourier transform. The purpose of the refinement is to find a model which best describes the intensities of the reflections. The core idea of the advanced refinement methods is to calculate atomic scattering factors from charge densities calculated by quantum mechanics. The basic assumption is that the charge density derived from theoretical calculations is in good agreement with the real density given in the crystal. A significant advantage of QCr methods is that they do not require high-resolution data (d < 0.5 Å) like in a conventional charge density analysis to obtain very precise information about the positions and displacements of the atoms in the crystal, especially of the light atoms. Nevertheless, it should be emphasized that high-resolution diffraction data are essential in order to obtain an accurate crystal structure using single-crystal X-ray diffraction.

For QCr refinement, a partitioning scheme is required, and two leading methods are discussed in this paper. In Hirshfeld atom refinement (HAR), the Hirshfeld partitioning scheme defines the electron density of an atom (Jayatilaka & Dittrich, 2008[Jayatilaka, D. & Dittrich, B. (2008). Acta Cryst. A64, 383-393.]; Capelli et al., 2014[Capelli, S. C., Bürgi, H.-B., Dittrich, B., Grabowsky, S. & Jayatilaka, D. (2014). IUCrJ, 1, 361-379.]). This method, which first appeared in the program TONTO (Jayatilaka & Grimwood, 2003[Jayatilaka, D. & Grimwood, D. J. (2003). Computational Science (ICCS 2003), edited by P. M. A. Sloot, D. Abramson, A. V. Bogdanov, Y. E. Gorbachev, J. J. Dongarra & A. Y. Zomaya, pp. 142-151. Berlin, Heidelberg: Springer.]) can nowadays be used in other structure refinement programs such as NoSpherA2/OLEX2 (Kleemiss et al., 2021[Kleemiss, F., Dolomanov, O. V., Bodensteiner, M., Peyerimhoff, N., Midgley, L., Bourhis, L. J., Genoni, A., Malaspina, L. A., Jayatilaka, D., Spencer, J. L., White, F., Grundkötter-Stock, B., Steinhauer, S., Lentz, D., Puschmann, H. & Grabowsky, S. (2021). Chem. Sci. 12, 1675-1692.]), DiSCaMb (Chodkiewicz et al., 2018[Chodkiewicz, M. L., Migacz, S., Rudnicki, W., Makal, A., Kalinowski, J. A., Moriarty, N. W., Grosse-Kunstleve, R. W., Afonine, P. V., Adams, P. D. & Dominiak, P. M. (2018). J. Appl. Cryst. 51, 193-199.]) or lamaGOET (Malaspina et al., 2021[Malaspina, L. A., Genoni, A. & Grabowsky, S. (2021). J. Appl. Cryst. 54, 987-995.]). Coordinates and anisotropic displacement parameters of hydrogen atoms can be determined for a number of examples with an accuracy that is otherwise only observed in neutron diffraction experiments (Ruth et al., 2022[Ruth, P. N., Herbst-Irmer, R. & Stalke, D. (2022). IUCrJ, 9, 286-297.]; Chodkiewicz et al., 2020[Chodkiewicz, M. L., Woińska, M. & Woźniak, K. (2020). IUCrJ, 7, 1199-1215.]; Woińska et al., 2021[Woińska, M., Chodkiewicz, M. L. & Woźniak, K. (2021). Chem. Commun. 57, 3652-3655.]). However, HAR is not yet a reliable substitute for neutron measurements in general, as it has been shown that it is still not possible to find suitable models for all systems and modifications are required (Chodkiewicz & Woźniak, 2025[Chodkiewicz, M. & Woźniak, K. (2025). IUCrJ, 12, 74-87.]). For this reason, intensive research is continuing on the method in order to optimize this approach. Work has been done in electron correlation handling. In principle, any theoretical method that differs in the treatment of electron correlation can be used to calculate a charge density (e.g. Hartree–Fock, DFT, MP2, CCSD). However, previous studies have already shown that the model charge density obtained from a DFT calculation is a good approximation for QCr refinement (Wieduwilt et al., 2020[Wieduwilt, E. K., Macetti, G., Malaspina, L. A., Jayatilaka, D., Grabowsky, S. & Genoni, A. (2020). J. Mol. Struct. 1209, 127934.]). DFT in HAR has been shown to differ significantly when the Hartree–Fock exchange is varied (Landeros-Rivera et al., 2023[Landeros-Rivera, B., Ramírez-Palma, D., Cortés-Guzmán, F., Dominiak, P. M. & Contreras-García, J. (2023). Phys. Chem. Chem. Phys. 25, 12702-12711.]). Another sophistication is to include the normal modes from a solid-state calculation in the refinement. In this procedure the frequencies of the modes are refined and a precise determination of the anisotropic displacement parameters is possible (Hoser & Madsen, 2016[Hoser, A. A. & Madsen, A. Ø. (2016). Acta Cryst. A72, 206-214.], Butkiewicz et al., 2025[Butkiewicz, H., Chodkiewicz, M., Madsen, A. Ø. & Hoser, A. A. (2025). IUCrJ, 12, 123-136.]). The influence of the partitioning scheme on the refinement result has been investigated and the advantages of alternative approaches have been analysed (Chodkiewicz et al., 2020[Chodkiewicz, M. L., Woińska, M. & Woźniak, K. (2020). IUCrJ, 7, 1199-1215.]; Chodkiewicz & Woźniak, 2025[Chodkiewicz, M. & Woźniak, K. (2025). IUCrJ, 12, 74-87.]). Until very recently, most HAR investigations use gas phase calculations with programs such as ORCA (Neese, 2022[Neese, F. (2022). WIREs Comput. Mol. Sci. 12, e1606.]), Gaussian (Frisch et al., 2016[Frisch, M. J., Trucks, G. W., Schlegel, H. B., Scuseria, G. E., Robb, M. A., Cheeseman, J. R., Scalmani, G., Barone, V., Petersson, G. A., Nakatsuji, H., Li, X., Caricato, M., Marenich, A. V., Bloino, J., Janesko, B. G., Gomperts, R., Mennucci, B., Hratchian, H. P., Ortiz, J. V., Izmaylov, A. F., Sonnenberg, J. L., Williams-Young, D., Ding, F., Lipparini, F., Egidi, F., Goings, J., Peng, B., Petrone, A., Henderson, T., Ranasinghe, D., Zakrzewski, V. G., Gao, J., Rega, N., Zheng, G., Liang, W., Hada, M., Ehara, M., Toyota, K., Fukuda, R., Hasegawa, J., Ishida, M., Nakajima, T., Honda, Y., Kitao, O., Nakai, H., Vreven, T., Throssell, K., Montgomery, J. A. Jr, Peralta, J. E., Ogliaro, F., Bearpark, M. J., Heyd, J. J., Brothers, E. N., Kudin, K. N., Staroverov, V. N., Keith, T. A., Kobayashi, R., Normand, J., Raghavachari, K., Rendell, A. P., Burant, J. C., Iyengar, S. S., Tomasi, J., Cossi, M., Millam, J. M., Klene, M., Adamo, C., Cammi, R., Ochterski, J. W., Martin, R. L., Morokuma, K., Farkas, O., Foresman, J. B. & Fox, D. J. (2016). Gaussian 16. Revision C.01. Gaussian Inc., Wallingford, Connecticut, USA.]), pySCF (Sun et al., 2020[Sun, Q., Zhang, X., Banerjee, S., Bao, P., Barbry, M., Blunt, N. S., Bogdanov, N. A., Booth, G. H., Chen, J., Cui, Z.-H., Eriksen, J. J., Gao, Y., Guo, S., Hermann, J., Hermes, M. R., Koh, K., Koval, P., Lehtola, S., Li, Z., Liu, J., Mardirossian, N., McClain, J. D., Motta, M., Mussard, B., Pham, H. Q., Pulkin, A., Purwanto, W., Robinson, P. J., Ronca, E., Sayfutyarova, E. R., Scheurer, M., Schurkus, H. F., Smith, J. E. T., Sun, C., Sun, S.-N., Upadhyay, S., Wagner, L. K., Wang, X., White, A., Whitfield, J. D., Williamson, M. J., Wouters, S., Yang, J., Yu, J. M., Zhu, T., Berkelbach, T. C., Sharma, S., Sokolov, A. Y. & Chan, G. K.-L. (2020). J. Chem. Phys. 153, 024109.]) or TONTO (Jayatilaka & Grimwood, 2003[Jayatilaka, D. & Grimwood, D. J. (2003). Computational Science (ICCS 2003), edited by P. M. A. Sloot, D. Abramson, A. V. Bogdanov, Y. E. Gorbachev, J. J. Dongarra & A. Y. Zomaya, pp. 142-151. Berlin, Heidelberg: Springer.]) being employed to calculate the aspherical atomic form factors. The importance of intermolecular interactions for the understanding of a solid in its physical and chemical properties is widespread in crystallography (Deringer et al., 2017[Deringer, V. L., George, J., Dronskowski, R. & Englert, U. (2017). Acc. Chem. Res. 50, 1231-1239.]) and is particularly emphasized in this work. In HAR applications, cluster charges and dipoles can be used to describe the effects of the environment on the charge density of a molecule (Woińska et al., 2014[Woińska, M., Jayatilaka, D., Spackman, M. A., Edwards, A. J., Dominiak, P. M., Woźniak, K., Nishibori, E., Sugimoto, K. & Grabowsky, S. (2014). Acta Cryst. A70, 483-498.]). An alternative to this HAR approach, which has so far been used scarcely, is the calculation under periodic boundary conditions (Ruth et al., 2022[Ruth, P. N., Herbst-Irmer, R. & Stalke, D. (2022). IUCrJ, 9, 286-297.]; Wall, 2016[Wall, M. E. (2016). IUCrJ, 3, 237-246.]). Recently, Dietmar Stalke's group presented a plane wave based method for HAR (PAW-HAR) and could show the benefit of periodic boundaries (Ruth et al., 2022[Ruth, P. N., Herbst-Irmer, R. & Stalke, D. (2022). IUCrJ, 9, 286-297.]).

In this work, we report an iterative refinement procedure which uses periodic boundary theoretical calculations with 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.]) to create a multipole description of the electron density for use in iterative least-squares refinement of X-ray diffraction data. We show that when using this method for D/L-serine and xylitol the hydrogen atoms in the hydrogen bonds are well described. In the case of xylitol, an improvement in the hydrogen atom positions compared with the gas phase HAR is observed using neutron diffraction results as a reference. This paper presents a method to independently compare the charge density from a solid-state calculation with the experimental charge density, thus allowing accurate comparisons between theory and experiment. For this purpose, the Hansen–Coppens multipole model is used, which has shown in numerous cases an inferior performance compared with the HAR approach in determining hydrogen atom positions (Chodkiewicz et al., 2024[Chodkiewicz, M., Patrikeev, L., Pawlędzio, S. & Woźniak, K. (2024). IUCrJ, 11, 249-259.]; Woińska et al., 2014[Woińska, M., Jayatilaka, D., Spackman, M. A., Edwards, A. J., Dominiak, P. M., Woźniak, K., Nishibori, E., Sugimoto, K. & Grabowsky, S. (2014). Acta Cryst. A70, 483-498.]). However, there is a weakness in the HAR with regards to the derivation of atomic displacement parameters compared with the multipole models (Köhler et al., 2019[Köhler, C., Lübben, J., Krause, L., Hoffmann, C., Herbst-Irmer, R. & Stalke, D. (2019). Acta Cryst. B75, 434-441.]).

The Hansen–Coppens multipole model, developed in the 1970s for crystallographic refinement, describes the aspherical charge density of an atom via spherical harmonics and radially adjustable functions (Hansen & Coppens, 1978[Hansen, N. K. & Coppens, P. (1978). Acta Cryst. A34, 909-921.]), leading to a description of the charge density via a set of multipole model parameters (MMPs). These MMPs can be determined directly from the diffraction experiment or from theoretically calculated structure factors. When fitted experimentally, this is referred to as a charge density study. Models have been developed using transferable MMPs derived from a theoretical calculation, the transferable aspherical atom model (TAAM), such as the Invariom model (Jha et al., 2020[Jha, K. K., Gruza, B., Kumar, P., Chodkiewicz, M. L. & Dominiak, P. M. (2020). Acta Cryst. B76, 296-306.]; Dittrich et al., 2004[Dittrich, B., Koritsánszky, T. & Luger, P. (2004). Angew. Chem. Int. Ed. 43, 2718-2721.]). However, there are certain limitations with the current multipole-based methods due to need for a reference to a database. This disadvantage can be circumvented by directly calculating the structure factors from the wavefunction. The CRYSTAL program offers the possibility of calculating dynamic and static structure factors from the wavefunction (Erba et al., 2013[Erba, A., Ferrabone, M., Orlando, R. & Dovesi, R. (2013). J. Comput. Chem. 34, 346-354.]). These can then be converted into theoretical MMPs (tMMPs). Such parameters can subsequently be used to refine the crystal structure, as in the TAAM or Invariom approach. By employing CRYSTAL to calculate the model it is possible to take into account the periodic effects on the charge density. Most recently, an application of TAAM was also developed to use multipole parameters generated directly from a solid-state calculation (Olech et al., 2024[Olech, B., Brázda, P., Palatinus, L. & Dominiak, P. M. (2024). IUCrJ, 11, 309-324.]). However, in this case a geometry from the literature is used and therefore information about the molecular structure is required.

2. Methodology

In order to coordinate all steps from the generation of the aspherical atomic form factors to the refinement and convergence test, a Python3 script was written, which has been given the name ReCrystal (refinement with CRYSTAL17). ReCrystal (Patzer, 2023[Patzer, M. (2023). PhD thesis. Bergische Universität Wuppertal, Germany.]) refers to a multipole model-based refinement in which optimized coordinates, harmonic oscillations and multipole parameters are generated with the help of periodic wavefunctions from the CRYSTAL17 program (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.]). At this point, note that ReCrystal does not claim to be an error-free refinement software, but rather serves as a tool for testing the approach. The ReCrystal program used for the project can be found on GitHub (https://github.com/MichaelPatzer/ReCrystal).

2.1. Generation of aspherical atom form factors

In the ReCrystal approach the atomic charge density in the crystal is described using the Hansen–Coppens multipole model [equation (1[link])]. Theoretical multipole parameters can be derived from theoretical static structure factors calculated directly by the CRYSTAL17 program. To evaluate the influence of resolution on the refinement indicators, ReCrystal was run at different resolutions (POB-TZVP/PBE). The main conclusion is that there is no significant change in the final R value and the resolution of the theoretical intensities in the range 0.9–0.3 Å (see next section). This observation can be explained by the fact that the multipole parameters in the main describe the valence electron density. The value of 0.8 Å has proven to be the most reliable resolution, so that an efficient convergence is found in the ReCrystal refinement resulting a low R1 value. Due to the fact that the results may differ slightly, we have decided to always specify the resolution used to calculate the tMMP in the ReCrystal refinement. Nevertheless, in order to obtain the tMMP with the highest possible accuracy and precision (low standard uncertainties in the refinement), the ratio of theoretical reflections to parameters should be as high as possible.

[\eqalignno { \rho_{\rm at}(r,\theta, \phi) = & \ P_c\rho_{\rm core}(r) + P_v\kappa^3 \rho_{\rm valence}(\kappa r) \cr & + \mathop \sum \limits_{l = 0}^{l_{\rm max}} \left [{\kappa'^3}R_l(\kappa'r) \mathop \sum \limits_{m = 0}^l P_{lm \pm}Y_{lm \pm}(\theta, \phi) \right]. & (1)}]

The remaining procedure is almost the same as for HAR (see Fig. 1[link]). The user of the ReCrystal software must select the basis set (e.g. POB-TZVP, def2-SVP, def2-TZVP) and a DFT functional before starting the refinement. In addition, the accuracy of the grid in the first Brillouin zone must be defined for the calculation of the k-points according to Pack–Monkhorst and Gilat. The resolution of the synthetic dataset used to obtain the tMMP is automatically set to the desired value. The multipole refinement is carried out with the program XD2006 (Volkov et al., 2006[Volkov, A. M. P., Ferrugia, L. J., Gatti, C., Mallinson, P. R., Richter, T. & Koritsanszky, T. (2006). XD2006. University at Buffalo, State University of New York, NY, USA; Università di Milano, Italy; University of Glasgow, UK; CNR-ISTM, Milano, Italy; and Middle Tennessee State University, TN, USA.]; https://xd.chem.buffalo.edu/). For the treatment of the radial components in the multipole model, the unpublished VM database contained in XD2006 is used. For this database, the individual exponents are derived from the density fit for each given orbital. The generation of the tMMP is done step by step to facilitate convergence. The weighting scheme, which is also implemented in SHELXL, is used for the refinement on F2 (see the supporting information, xd.mas). In the first part, the multipole occupations of all atoms are refined up to the octupole level. In the second step, the kappa parameters are refined individually for each atom and the multipole occupations are fixed. Finally, the kappa parameters and all multipole parameters are refined together. The multipoles of the hydrogen atoms are fixed in the third step to reduce correlation effects. The scale factor is refined in all steps of the least-squares procedure.

[Figure 1]
Figure 1
Comparison of the HAR (top) and ReCrystal (bottom) procedures. The methods differ mainly in the treatment of the aspherical atomic charge density. The former method uses Hirshfeld atoms. The latter method uses multipole parameters from the Hansen–Coppens multipole model. (1) Generation of a (periodic) wavefunction; (2) generation of aspherical atomic form factors; (3) refinement of x, y, z and ADPs with fixed theoretical aspherical atomic form factors; and (4) convergence test: negative: calculation of new periodic wavefunction using the refined (x, y, z) values, positive: end of refinement.

2.2. Crystal structure refinement

In this step, the objective is to refine the three positional coordinates x, y and z and the six anisotropic displacement parameters of the atoms in the asymmetric unit. The theoretical multipole parameters are included in the refinement with fixed values and describe the static charge density in the crystal. The experimental diffraction intensities are used as a reference in the refinement. The weighting scheme, which is also implemented in SHELXL, is used for the refinement on F2 (see the supporting information, xd.mas). The dispersion correction is switched on again in this step and the tabulated element-specific values implemented in XD2006 are used for f′ and f′′ (default Mo Kα radiation). The entire refinement is again executed in three steps. In the first step, the coordinates and harmonically anisotropic displacements of all atoms heavier than hydrogen are refined. The coordinates and displacement parameters of the hydrogen atoms are not refined in this step. In the second step, the coordinates of the hydrogen atoms are added to the refinement and the displacement parameters are included as isotropic. Finally, all coordinates are freely refined and the displacements of all atoms are treated anisotropically. This three-step procedure is intended to counteract convergence problems due to high correlations. The scale factor is refined in all least-squares procedures. The refined coordinates are checked for convergence in the next step.

For checking convergence, the current version of ReCrystal uses the parameter φdiff [equation (2[link])]. This value checks the geometrically averaged change of coordinates in the entire asymmetric unit after a refinement step. If the value of φdiff is less than 1 × 10−4 Å, the refinement is considered to be finished. This means that the coordinates have changed by less than 0.1 mÅ on average within the refinement cycle. Convergence is controlled by the value of φdiff in the ReCrystal program from the second cycle onwards.

[\varphi_{\rm diff} = {1 \over {{N_{\rm asym}}}} \sqrt {\mathop \sum \limits_{n = 1}^{N_{\rm asym}} {{\left({\sqrt {{x_{n,{\rm diff}}}^2 + {y_{n,{\rm diff}}}^2 + {z_{n,{\rm diff}}}^2} } \right)}^2}}. \eqno (2)]

In addition, the atom-specific mean-squared-displacement amplitude for the coordinates and anisotropic displacement parameters are calculated and documented in the output files. If the convergence criterion is not fulfilled, a new periodic wavefunction is computed using the new refined coordinates. The refinement process is repeated (Fig. 2[link]).

[Figure 2]
Figure 2
Detailed representation of the sequence of a ReCrystal refinement. Initial parameters must be defined once; (1) generation of a periodic wavefunction (2) generation of static structure factors and tMMPs; (3) refinement of x, y, z and ADPs with fixed theoretical MMPs; (4) convergence test: negative, then a new periodic wavefunction is calculated using the refined x, y, z values; positive: end of ReCrystal refinement; Ncyc: cycle number; Nmax: maximum of Ncyc defined by user.

3. D/L-serine: basis set and DFT-functional dependence

A key issue to be addressed is the influence of the basis set and the DFT functional on the ReCrystal refinement result. For this purpose, the X-ray diffraction 100 K dataset of D/L-serine by Peter Luger's group (Dittrich et al., 2005[Dittrich, B., Hübschle, C. B., Messerschmidt, M., Kalinowski, R., Girnt, D. & Luger, P. (2005). Acta Cryst. A61, 314-320.]) was refined with various basis sets and functionals (Table 1[link], CCDC No. 273585). Note that in some cases there were convergence problems on the part of CRYSTAL17 or XD, so that not all combinations in Table 1[link] can be assigned an R value. XD was unable to generate the theoretical multipole parameters from a charge density calculated with a small basis set. This includes the MINIX and STO 3G basis sets. It is assumed that especially the high-angle reflections, which can be assigned to the charge density close to the nucleus, deviate strongly from the expected value due to the small size of the Gaussian basis set. As a result, due to the inadequate description of the near-nuclear charge density, the convergence in the program XD is not fulfilled when generating the multipole parameters in ReCrystal. When combining the basis set def2-TZVP with the hybrid functionals, such as B3LYP, wb97 or M062X, no convergence was achieved in CRYSTAL17-SCF. The combination of the PBE and BLYP functional and the def2-TZVP basis set provides the best R1 value of 1.87%. Refining the structure with tMMP generated with intensities up to a resolution of 0.8 Å resulted in an R1 of 1.82%. After testing resolutions to generate tMMP between 0.3 and 0.9 Å, the best R1 was found at a resolution of 0.8 Å. In view of this variation, we suggest that the resolution to which the theoretical calculations were undertaken in the ReCrystal refinement should be given [ReCrystal/(number) Å]. Thus, for example, (ReCrystal/0.8 Å). Additionally, it is recommended to test a ReCrystal refinement with the Ahlrich basis set def2-TZVP, but keep in mind that the convergence in CRYSTAL17 might not be fulfilled due to the linear dependencies. The crystal structure was refined with both ReCrystal/0.8 Å and NoSpherA2/HAR and no significant differences can be recognized between the atomic displacement ellipsoid plots, except perhaps for the directions of the principal axes of the atomic displacement ellipsoids for the hydrogen atoms (Fig. 3[link]). In the HAR, the ellipsoid of the hydrogen atom H3 is elongated in the direction of the hydrogen bond compared with ReCrystal/0.8 Å. A comparison with neutron diffraction results provides additional insight into the advantage of the method (see Section 5[link] on xylitol). The 0.8 Å value refers exclusively to the resolution of the synthetic dataset of diffraction intensities used to determine the theoretical multipole parameters. The experimental dataset is not truncated and its full resolution is used to determine the displacement parameters and atomic coordinates.

Table 1
Dependence of R1/% and wR2/% in the ReCrystal/0.4 Å refinement on the choice of the basis set and DFT functional; refinement of D/L-serine (rac.) ten cycles

A dash refers to convergence not fulfilled in CRYSTAL17.

  PBE BLYP B3LYP CAM-B3LYP wB97 M062X
R1 wR2 R1 wR2 R1 wR2 R1 wR2 R1 wR2 R1 wR2
POB-DZVP 1.99 4.61 1.98 4.59 1.96 4.56 2.04 4.76 1.97 4.58
6-31G 1.93 4.45 1.92 4.43 1.92 4.42 1.92 4.42 1.93 4.43 1.92 4.42
def2-SVP 1.93 4.47 1.92 4.44 1.92 4.43 1.92 4.42 1.92 4.43 1.93 4.45
POB-DZVPP 1.92 4.46 1.91 4.43
POB-TZVP 1.91 4.45 1.91 4.44 1.90 4.41 1.89 4.39 1.89 4.41 1.90 4.41
def2-TZVP 1.87 4.36 1.87 4.36
[Figure 3]
Figure 3
Atomic displacement ellipsoid plot for refinement with ReCrystal/0.8 Å (top), NoSphereA2 [ORCA (version 5.0.3); bottom] with the basis set def2-TZVP and functional PBE; ellipsoid representation at the 50% probability level (hydrogen – white, carbon – grey, nitro­gen – blue, oxygen – red).

Table 1[link] reveals that the result with the lowest R1 value is obtained by refinement with the def2-TZVP basis set. ReCrystal must be analysed for the behaviour of the refinement and the defined resolution in order to calculate theoretical intensities. The result that fits best is obtained at 0.8 Å (Fig. 4[link]). This result is explained by the fact that weaker diffraction intensities at resolution values of less than 0.8 Å lead to numerical inaccuracies. In addition, the ReCrystal method converges faster at 0.8 Å after seven cycles starting with the IAM geometry and the obtained coordinates are suitable (see xylitol, Section 5[link]).

[Figure 4]
Figure 4
Comparison of the result for the ReCrystal refinement at different resolutions for the calculation of theoretical diffraction intensities.

The results of the refinement with different basis sets can be compared using the residual density analysis according to Meindl & Henn (2008[Meindl, K. & Henn, J. (2008). Acta Cryst. A64, 404-418.]). Fig. 5[link] shows that the Ahlrichs basis set def2-TZVP provides a more suitable charge density model than the calculation with POB-TZVP and POB-DZVP.

[Figure 5]
Figure 5
Residual density plot for the ReCrystal/0.4 Å refinement of serine for three basis sets: POB-DZVP and POB-TZVP compared with def2-TZVP.

4. Bond distances from ReCrystal compared with HAR

The periodic boundary condition used in ReCrystal does not lead to any significant difference in the bond distances compared with the refinement using aspherical atomic form factors from a gas phase calculation with HAR. It must be emphasized at this point, however, that ReCrystal is based on multipole model atoms and HAR on Hirshfeld atoms. These fundamentally different approaches have significant differences. These include the fact that ReCrystal does not use the entire information of the wavefunction and the MMPs cannot exactly reproduce the calculated wavefunction (Bytheway et al., 2002[Bytheway, I., Chandler, G. S. & Figgis, B. N. (2002). Acta Cryst. A58, 451-459.]).

Both HAR and ReCrystal significantly increase the bond distances to the hydrogen atoms compared with the IAM. This is due to the consideration of the bonding electron density. The bond density is relatively high compared with the total density of the hydrogen atom, so the effect is particularly noticeable for hydrogen. Thus, the hydrogen atom does not have core electrons that can be reasonably described by the IAM (Fig. 6[link]) (Stewart et al., 1965[Stewart, R. F., Davidson, E. R. & Simpson, W. T. (1965). J. Chem. Phys. 42, 3175-3187.]) The next step is to compare these results with neutron diffraction. Due to the lack of a neutron diffraction result for D/L-serine, another molecular crystal will be discussed in the next section.

[Figure 6]
Figure 6
Comparison of bond distances in D/L-serine for the refinement with IAM (SHELXL), HAR [NoSpherA2, ORCA (version 5.0.3), def2-TZVP/PBE] and ReCrystal/0.8 Å (XD, CRYSTAL17, def2-TZVP/PBE).

5. Comparison to neutron diffraction

The molecular crystal of the achiral organic compound xylitol crystallizes in the Sohncke space group P212121. The crystal structure was first published by Kim & Jeffrey (1969[Kim, H. S. & Jeffrey, G. A. (1969). Acta Cryst. B25, 2607-2613.]), who were able to obtain an X-ray diffraction dataset with Cu Kα radiation. This work is based on a high-resolution dataset (dmin = 0.41 Å), the dataset for X-ray diffraction with synchrotron radiation at 122 K (CCDC No. 1432562) already published by Madsen et al. (2004[Madsen, A. Ø., Sørensen, H. O., Flensburg, C., Stewart, R. F. & Larsen, S. (2004). Acta Cryst. A60, 550-561.]). The structure from neutron diffraction at 122 K published by Madsen et al. (2003[Madsen, A. Ø., Mason, S. & Larsen, S. (2003). Acta Cryst. B59, 653-663.]) (CCDC No. 223330) is also used as a reference (Madsen et al., 2003[Madsen, A. Ø., Mason, S. & Larsen, S. (2003). Acta Cryst. B59, 653-663.]). A pronounced network of intermolecular hydrogen bonds is apparent in the crystal structure of the compound. After optimizing the ReCrystal procedure, using the basis set def2-TZVP/PBE and calculating tMMP with theoretical intensities up to 0.8 Å, the full performance of ReCrystal/0.8 Å is tested if one can obtain reasonable agreement to neutron diffraction bond distances. An extinction correction is used additionally in the iterative refinement, which leads to an improvement in the residual electron density and the R values. The extinction correction was applied in each iteration of NoSpherA2/HAR and ReCrystal. It was observed that the bond distances are shortened by a few mÅ by introducing the extinction correction (Fig. S6 of the supporting information compared with Fig. 8). Visually, the difference between gas phase HAR and ReCrystal/0.8 Å is not obvious from the perspective of the atomic displacement ellipsoid plot (Fig. 7[link]). Differences can only be found in the shape of the hydrogen ellipsoids bound to oxygen atoms.

[Figure 7]
Figure 7
Atomic displacement ellipsoid plot of xylitol for (top) ReCrystal/0.8 Å (def2-TZVP/PBE) and (bottom) HAR [NoSpherA2/def2-TZVP/PBE/ORCA (version 5.0.3)]; ellipsoid representation at the 50% probability level.

An improvement compared with gas phase HAR (NoSpherA2/def2-TZVP/PBE) can be observed for the determination of the C—H and O—H bond distances (see Fig. 8[link]). The C—H bond distances can be determined on average with an agreement to neutron diffraction of less than 20 mÅ difference. This example shows that a reasonable determination of the coordinates is possible with ReCrystal even though the computational effort is enlarged. In addition, the multipoles can be used for a further refinement of the charge density as shown for serine. Inspection of the fractal dimension analysis of the electron residual density, based on 9790 Bragg intensities, shows that the gas phase HAR is in close agreement with the ReCrystal result (Figs. S2 and S3). In addition, it is demonstrated that the anisotropic displacement parameters, especially those of the hydrogen atoms bound to oxygen, are systematically more accurate with ReCrystal than with a HAR based on gas phase calculations (Fig. 9[link]). The interaction density indicates that the polarization is pronounced due to hydrogen bonding on the hydrogen atoms bound to the oxygen (Fig. S5).

[Figure 8]
Figure 8
Box-whisker plot analysis showing the difference of bond distances between neutron diffraction and the model for single-crystal X-ray diffraction for xylitol: HAR [NoSpherA2/ORCA (version 5.0.3)/def2-TZVP/PBE], ReCrystal/0.8 Å (def2-TZVP/PBE) and IAM (SHELXL).
[Figure 9]
Figure 9
Box-whisker plot analysis showing the difference in the ADPs between neutron diffraction and the model for single-crystal X-ray diffraction for xylitol: HAR [NoSpherA2/ORCA (version 5.0.3)/def2-TZVP/PBE] and ReCrystal/0.8 Å (def2-TZVP/PBE).

6. Influence of environmental effects on the molecular charge density

From a theoretical point of view, a difference in the charge density of the isolated molecule and the molecule in the crystal can be expected due to the intermolecular interactions (Spackman et al., 1999[Spackman, M. A., Byrom, P. G., Alfredsson, M. & Hermansson, K. (1999). Acta Cryst. A55, 30-47.]). This can be shown in the theoretical difference Fourier map from structure factors generated from isolated molecular densities and interacting molecules (Fig. 10[link]). The calculated interaction density, which is taken into account when periodic boundary conditions are used in the wavefunction calculation, is illustrated.

[Figure 10]
Figure 10
Difference Fourier map for refinement of a synthetic dataset of serine (CRYSTAL17 def2-TZVP/PBE, dmin = 0.55 Å) with HAR [isolated molecule, NoSpherA2, ORCA (version 5.0), def2-TZVP/PBE] to illustrate the interaction density, isolevel 0.04 e Å−3.

The comparison between the R1all of the refinement with an isolated charge density from an ORCA (version 5.0.3) calculation and the refinement with the periodic density from CRYSTAL17 shows no significant differences. The R1all is 2.28% with HAR and 2.35% with ReCrystal/0.4 Å when using the basis set def2-TZVP and the DFT functional PBE. ReCrystal/0.8 Å results an R1all of 2.23%. The ORTEP plots from HAR and ReCrystal show only slight differences in the displacement ellipsoids of the hydrogen atoms. Otherwise, no significant structural differences can be observed (Figs. 3[link] and 6[link]).

For serine, an improvement in the fractal analysis of the residual density compared with the HAR of the isolated molecule is not evident (see Fig. 11[link]). The effects of environmental influences on the charge density are not evident. The fractal analysis based on 4930 independent reflections yields the same integrated residual electron density [\rho_{\rm gross}] of approximately 11 e Å−3. This example is mainly used to test the ReCrystal process and to get a general impression of the refinement. It will be necessary to study more systems in order to establish the influence of the charge density by the molecules in the environment that can be observed in an experiment, particularly in the case of extended systems such as coordination polymers and covalent structures.

[Figure 11]
Figure 11
Henn–Meindl residual density plot based on 4930 independent reflections of refinement with ReCrystal/0.8 Å, HAR/NoSpherA2 with the basis set def2-TZVP, and functional PBE and IAM for the D/L-serine dataset.

7. Charge density analysis

Here we present an advantageous aspect of the Hansen–Coppens model based on ReCrystal tMMP in charge density analysis. In particular, ReCrystal can be used to analyse weak interactions such as hydrogen bonds. The D/L-serine molecular crystal has been chosen in this case mainly because of the weak molecular interactions in the form of hydrogen bonds in the crystal. Qualitatively, the representation of the deformation density and the representation of the bond critical points (BCPs) can provide an initial impression of the quality of the refinement model (Fig. 12[link]).

[Figure 12]
Figure 12
Electron density from the difference between the structure factors of the aspherical model and the spherical model for ReCrystal (left: CRYSTAL17, def2-TZVP, PBE) and the multipole model based on tMMP from ReCrystal/0.8 Å, contour +0.3 e Å−3 (ellipsoids represented at the 50% probability level); right: QTAIM analysis for BCPs (red) of the tMMP with ReCrystal.

Fig. 10[link] shows a positive interaction density in the carbon–nitro­gen bond due to interaction effects. This interaction effect is taken into account in the ReCrystal refinement. If the density at the (3,−1) BCP is determined after gas phase HAR for the wavefunction [def2-TZVP/PBE; Multiwfn (version 3.6; Lu & Chen, 2012[Lu, T. & Chen, F. (2012). J. Comput. Chem. 33, 580-592.])], the density is 1.564 e Å−3. As expected, the density with ReCrystal is higher and is 1.659 e Å−3 with the same basis set and functional. The expectation can be additionally checked with TOPOND17 using the CRYSTAL17 calculated wavefunction on the POB-TZVP basis level. The value is 1.613 e Å−3. This shows that the interaction effects are taken into account in the ReCrystal refinement. The interaction density obtained from multipole refinement based on the tMMP of ReCrystal converges at 1.626 e Å−3.

The next step is to investigate whether the tMMP of the ReCrystal refinement can represent the weak interaction effects of the CRYSTAL17 wavefunction. To check this, the results of the QTAIM analysis are compared for the weak hydrogen-bonding densities (Fig. 13[link] and Table 2[link]). The ReCrystal/0.4 Å tMMP from a calculation def2-TZVP/PBE is analysed with TOPXD for the last refinement cycle. The CRYSTAL17 wavefunction cannot be analysed with TOPOND17 for the last refinement cycle because TOPOND17 is not able to analyse F functions. This problem is overcome in the latest version of TOPOND23, which has not been accessible for us until now. For this reason, the tMMPs from the POB-TZVP/PBE refinement are compared with the CRYSTAL17 wavefunction. From this point of view one can observe that the values for the charge density can be represented in good agreement with a maximum change of 0.01 e Å−3. This value is acceptable with respect to the fact that the interaction density effect occurs in a range slightly higher than 0.01 e Å−3. The Laplacian differs slightly compared with the theoretical calculations. This observation has already been described in so far that the multipole model sometimes has problems representing the Laplacian for weak interactions in general (Bytheway et al., 2002[Bytheway, I., Chandler, G. S. & Figgis, B. N. (2002). Acta Cryst. A58, 451-459.]).

Table 2
BCPs in the crystal structure of D/L-serine for intermolecular bonds

First row: free multipole refinement (MM, XD2016, VM database, TOPXD); second and third row: ReCrystal/0.4 Å (PBE, TOPXD); fourth row: TOPOND from the CRYSTAL17 wavefunction (POB: POB-TZVP/PBE; def2: def2-TZVP/PBE).

  Bond D (Å) D1 (Å) D2 (Å) CP ρCP (e Å−3) 2ρCP (e Å−5) λ1 (e Å−5) λ2 (e Å−5) λ3 (e Å−5) ε
1 MM 1.823 0.638 1.185 (3,−1) 0.228 2.142 −1.530 −1.255 4.928 0.219
ReCrystal def2 1.819 0.649 1.170 (3,−1) 0.252 2.040 −1.533 −1.447 5.020 0.060
ReCrystal POB 1.814 0.639 1.175 (3,−1) 0.248 2.207 −1.513 −1.388 5.108 0.090
TOPOND POB   (3,−1) 0.249 2.482 −1.398 −1.350 5.229 0.026
2 MM 1.793 0.635 1.158 (3,−1) 0.264 2.766 −1.377 −1.250 5.393 0.101
ReCrystal def2 1.794 0.648 1.146 (3,−1) 0.267 2.710 −1.474 −1.398 5.582 0.054
ReCrystal POB 1.793 0.641 1.152 (3,−1) 0.265 2.745 −1.458 −1.364 5.567 0.069
TOPOND POB   (3,−1) 0.250 2.940 −1.446 −1.350 5.736 0.071
3 MM 1.805 0.638 1.167 (3,−1) 0.248 2.174 −1.653 −1.263 5.090 0.309
ReCrystal def2 1.798 0.643 1.155 (3,−1) 0.256 2.405 −1.556 −1.416 5.377 0.099
ReCrystal POB 1.823 0.653 1.170 (3,−1) 0.239 2.465 −1.381 −1.253 5.100 0.102
TOPOND POB   (3,−1) 0.229 2.771 −1.253 −1.205 5.229 0.037
4 MM 1.731 0.596 1.135 (3,−1) 0.335 1.439 −2.329 −2.191 5.959 0.063
ReCrystal def2 1.726 0.599 1.127 (3,−1) 0.323 1.767 −2.121 −2.079 5.967 0.021
ReCrystal POB 1.720 0.586 1.134 (3,−1) 0.302 2.204 −1.996 −1.916 6.115 0.042
TOPOND POB   (3,−1) 0.290 3.181 −1.687 −1.663 6.531 0.024
[Figure 13]
Figure 13
Assignment of hydrogen bonds analysed by QTAIM for verification of the use of the ReCrystal tMMP.

In summary, ReCrystal refinement with tMMP is a reproduction of the CRYSTAL17 wavefunction, but with a minor gap of quantitative information. There are several reasons for this situation. One reason is that intensities close to 0 have to be calculated in reciprocal space for the full transformation of the real-space charge density into Bragg intensities, which is generally limited. For this example, ReCrystal uses intensities in reciprocal space up to 0.4 Å. Another point is that the MMPs cannot reproduce the charge density exactly because of the use of a tabulated form of exponential coefficients for the radial part. When fitting the theoretical intensity values, R1 drops to small values slightly greater than 0%. Nevertheless, the tMMP serves as a reasonable basis for charge density analysis, especially if one is interested in weak interaction effects.

8. Outlook and discussion

Multipole model-based iterative refinement (Mbit refinement) is revealed as a new advanced and generally applicable tool in QCr for the refinement of atomic positions and nuclear vibrations. The implementation of iterative refinement with theoretical multipole parameters was presented in the ReCrystal procedure (`Refinement with CRYSTAL17'). The resulting coordinates were shown to be comparable to those obtained by HAR. An even better performance than with the gas phase HAR has been observed. Further analysis of the wide range of molecular crystals will be the focus of future research.

Refined multipole parameters in a multipole refinement can account for effects that are not related to bonding electron density [e.g. disorder, anharmonic motion (Mallinson et al., 1988[Mallinson, P. R., Koritsanszky, T., Elkaim, E., Li, N. & Coppens, P. (1988). Acta Cryst. A44, 336-343.])]. Note that crystals are not ideal objects. Rather, crystals are imperfect systems with dislocations, vacancies and interstitial atoms that affect the diffraction pattern. The major concern is overinterpretation of these crystal construction errors and experimental errors on the observed intensities. Crystal defects can produce undesirable systematic errors in a multipole refinement. From the perspective of experimental charge density investigation, ideally one wants to determine the charge density from the diffraction pattern in order to investigate the true bonding situation. At this point, theory can contribute decisively to the study of systems of interest. Theoretical multipole parameters, which can be generated in the ReCrystal code, generate parameters that are already optimized and avoid the risk of describing defects caused by non-ideal properties of a crystal in a charge density study using the conventional multipole model. ReCrystal thus represents a modern approach to the study of charge density, combining the best of both experimental and theoretical techniques.

However, there are limitations to overcome. One problem arises when refining atoms on special positions. Because of the use of XD2006, special care must be taken when processing atoms on special positions. This problem no longer exists in XD2024, so this program will have to be incorporated into ReCrystal in future. In addition, refinement with ReCrystal and CRYSTAL17 is more computationally intensive than simple HAR. This is mainly due to the fact that after calculating the periodic wavefunction, a synthetic dataset must be created from which the aspherical atom form factors are then generated.

The combination of the ReCrystal refinement with a preceding HAR has proven to be particularly advantageous (Fig. 14[link]). Starting from the HAR structure, the ReCrystal refinement converges faster to the minimum and reliably produces a charge density model that can reduce the residual charge density compared with IAM. It should also be possible to combine the ReCrystal refinement with the normal mode refinement developed by Hoser & Madsen (2016[Hoser, A. A. & Madsen, A. Ø. (2016). Acta Cryst. A72, 206-214.]).

[Figure 14]
Figure 14
Optional flowchart for ReCrystal refinement of molecular crystals; ReCrystal also provides access to experimental charge density investigations on a theoretical basis with optimized coordinates and displacement parameters.

So far, this method has been tested intensively on molecular systems (Patzer, 2023[Patzer, M. (2023). PhD thesis. Bergische Universität Wuppertal, Germany.]). Note that the ReCrystal code has since been the subject of optimization and testing. This method can be used in particular for systems whose binding motifs are periodic in nature such as hydrogen-bonding motifs, coordination polymers and covalent structures. This is the first step for future research to explore the full periodic charge density in the crystalline solid state with the newly developed method based on periodic boundary calculations. It has already been established from a theoretical perspective that the ReCrystal process also works for these periodic systems to refine hydrogen atom positions, particularly those involved in intermolecular hydrogen bonds. This has been clearly demonstrated by comparisons with the results of neutron diffraction.

9. Related literature

The following reference is cited in the supporting information: Larsen (1985[Larsen, R. D. (1985). J. Chem. Educ. 62, 302-305.]).

Supporting information


Acknowledgements

Michael Patzer would like to thank Richard Goddard for the fruitful and supportive discussions on this project. Michael Patzer would also like to thank the members of the department of Chemical Crystallography and Electron Microscopy for their support and the Max-Planck-Institut für Kohlenforschung for providing the infrastructure for the research.

Conflict of interest

We declare that there are no conflicts of interest.

References

First citationButkiewicz, H., Chodkiewicz, M., Madsen, A. Ø. & Hoser, A. A. (2025). IUCrJ, 12, 123–136.  CrossRef CAS PubMed IUCr Journals Google Scholar
First citationBytheway, I., Chandler, G. S. & Figgis, B. N. (2002). Acta Cryst. A58, 451–459.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationCapelli, S. C., Bürgi, H.-B., Dittrich, B., Grabowsky, S. & Jayatilaka, D. (2014). IUCrJ, 1, 361–379.  Web of Science CSD CrossRef CAS PubMed IUCr Journals Google Scholar
First citationChodkiewicz, M., Patrikeev, L., Pawlędzio, S. & Woźniak, K. (2024). IUCrJ, 11, 249–259.  Web of Science CrossRef CAS PubMed IUCr Journals Google Scholar
First citationChodkiewicz, M. & Woźniak, K. (2025). IUCrJ, 12, 74–87.  Web of Science CSD CrossRef CAS PubMed IUCr Journals Google Scholar
First citationChodkiewicz, M. L., Migacz, S., Rudnicki, W., Makal, A., Kalinowski, J. A., Moriarty, N. W., Grosse-Kunstleve, R. W., Afonine, P. V., Adams, P. D. & Dominiak, P. M. (2018). J. Appl. Cryst. 51, 193–199.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationChodkiewicz, M. L., Woińska, M. & Woźniak, K. (2020). IUCrJ, 7, 1199–1215.  Web of Science CSD CrossRef CAS PubMed IUCr Journals Google Scholar
First citationDeringer, V. L., George, J., Dronskowski, R. & Englert, U. (2017). Acc. Chem. Res. 50, 1231–1239.  Web of Science CrossRef CAS PubMed Google Scholar
First citationDittrich, B., Hübschle, C. B., Messerschmidt, M., Kalinowski, R., Girnt, D. & Luger, P. (2005). Acta Cryst. A61, 314–320.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationDittrich, B., Koritsánszky, T. & Luger, P. (2004). Angew. Chem. Int. Ed. 43, 2718–2721.  Web of Science CSD CrossRef CAS Google Scholar
First citationDittrich, B., Strumpel, M., Schäfer, M., Spackman, M. A. & Koritsánszky, T. (2006). Acta Cryst. A62, 217–223.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationDominiak, P. M., Volkov, A., Li, X., Messerschmidt, M. & Coppens, P. (2007). J. Chem. Theory Comput. 3, 232–247.  Web of Science CrossRef CAS PubMed 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 citationErba, A., Ferrabone, M., Orlando, R. & Dovesi, R. (2013). J. Comput. Chem. 34, 346–354.  Web of Science CrossRef CAS PubMed Google Scholar
First citationFrisch, M. J., Trucks, G. W., Schlegel, H. B., Scuseria, G. E., Robb, M. A., Cheeseman, J. R., Scalmani, G., Barone, V., Petersson, G. A., Nakatsuji, H., Li, X., Caricato, M., Marenich, A. V., Bloino, J., Janesko, B. G., Gomperts, R., Mennucci, B., Hratchian, H. P., Ortiz, J. V., Izmaylov, A. F., Sonnenberg, J. L., Williams-Young, D., Ding, F., Lipparini, F., Egidi, F., Goings, J., Peng, B., Petrone, A., Henderson, T., Ranasinghe, D., Zakrzewski, V. G., Gao, J., Rega, N., Zheng, G., Liang, W., Hada, M., Ehara, M., Toyota, K., Fukuda, R., Hasegawa, J., Ishida, M., Nakajima, T., Honda, Y., Kitao, O., Nakai, H., Vreven, T., Throssell, K., Montgomery, J. A. Jr, Peralta, J. E., Ogliaro, F., Bearpark, M. J., Heyd, J. J., Brothers, E. N., Kudin, K. N., Staroverov, V. N., Keith, T. A., Kobayashi, R., Normand, J., Raghavachari, K., Rendell, A. P., Burant, J. C., Iyengar, S. S., Tomasi, J., Cossi, M., Millam, J. M., Klene, M., Adamo, C., Cammi, R., Ochterski, J. W., Martin, R. L., Morokuma, K., Farkas, O., Foresman, J. B. & Fox, D. J. (2016). Gaussian 16. Revision C.01. Gaussian Inc., Wallingford, Connecticut, USA.  Google Scholar
First citationGenoni, A., Bučinský, L., Claiser, N., Contreras–García, J., Dittrich, B., Dominiak, P. M., Espinosa, E., Gatti, C., Giannozzi, P., Gillet, J.-M., Jayatilaka, D., Macchi, P., Madsen, A. Ø., Massa, L., Matta, C. F., Merz, K. M. Jr, Nakashima, P. N. H., Ott, H., Ryde, U., Schwarz, K., Sierka, M. & Grabowsky, S. (2018). Chem. A Eur. J. 24, 10881–10905.  CrossRef CAS Google Scholar
First citationGrabowsky, S., Genoni, A. & Bürgi, H.-B. (2017). Chem. Sci. 8, 4159–4176.  Web of Science CrossRef CAS PubMed Google Scholar
First citationHansen, N. K. & Coppens, P. (1978). Acta Cryst. A34, 909–921.  CrossRef CAS IUCr Journals Web of Science Google Scholar
First citationHoser, A. A. & Madsen, A. Ø. (2016). Acta Cryst. A72, 206–214.  Web of Science CSD CrossRef IUCr Journals Google Scholar
First citationJayatilaka, D. & Dittrich, B. (2008). Acta Cryst. A64, 383–393.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationJayatilaka, D. & Grimwood, D. J. (2003). Computational Science (ICCS 2003), edited by P. M. A. Sloot, D. Abramson, A. V. Bogdanov, Y. E. Gorbachev, J. J. Dongarra & A. Y. Zomaya, pp. 142–151. Berlin, Heidelberg: Springer.  Google Scholar
First citationJha, K. K., Gruza, B., Kumar, P., Chodkiewicz, M. L. & Dominiak, P. M. (2020). Acta Cryst. B76, 296–306.  Web of Science CSD CrossRef IUCr Journals Google Scholar
First citationKim, H. S. & Jeffrey, G. A. (1969). Acta Cryst. B25, 2607–2613.  CSD CrossRef CAS IUCr Journals Web of Science Google Scholar
First citationKleemiss, F., Dolomanov, O. V., Bodensteiner, M., Peyerimhoff, N., Midgley, L., Bourhis, L. J., Genoni, A., Malaspina, L. A., Jayatilaka, D., Spencer, J. L., White, F., Grundkötter-Stock, B., Steinhauer, S., Lentz, D., Puschmann, H. & Grabowsky, S. (2021). Chem. Sci. 12, 1675–1692.  Web of Science CSD CrossRef CAS Google Scholar
First citationKöhler, C., Lübben, J., Krause, L., Hoffmann, C., Herbst-Irmer, R. & Stalke, D. (2019). Acta Cryst. B75, 434–441.  Web of Science CSD CrossRef IUCr Journals Google Scholar
First citationLanderos-Rivera, B., Ramírez-Palma, D., Cortés-Guzmán, F., Dominiak, P. M. & Contreras-García, J. (2023). Phys. Chem. Chem. Phys. 25, 12702–12711.  Web of Science CAS PubMed Google Scholar
First citationLarsen, R. D. (1985). J. Chem. Educ. 62, 302–305.  CrossRef CAS Google Scholar
First citationLu, T. & Chen, F. (2012). J. Comput. Chem. 33, 580–592.  Web of Science CrossRef PubMed Google Scholar
First citationMadsen, A. Ø., Mason, S. & Larsen, S. (2003). Acta Cryst. B59, 653–663.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationMadsen, A. Ø., Sørensen, H. O., Flensburg, C., Stewart, R. F. & Larsen, S. (2004). Acta Cryst. A60, 550–561.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationMalaspina, L. A., Genoni, A. & Grabowsky, S. (2021). J. Appl. Cryst. 54, 987–995.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationMallinson, P. R., Koritsanszky, T., Elkaim, E., Li, N. & Coppens, P. (1988). Acta Cryst. A44, 336–343.  CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationMeindl, K. & Henn, J. (2008). Acta Cryst. A64, 404–418.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationNeese, F. (2022). WIREs Comput. Mol. Sci. 12, e1606.  Google Scholar
First citationOlech, B., Brázda, P., Palatinus, L. & Dominiak, P. M. (2024). IUCrJ, 11, 309–324.  Web of Science CSD CrossRef CAS PubMed IUCr Journals Google Scholar
First citationPatzer, M. (2023). PhD thesis. Bergische Universität Wuppertal, Germany.  Google Scholar
First citationRuth, P. N., Herbst-Irmer, R. & Stalke, D. (2022). IUCrJ, 9, 286–297.  Web of Science CSD CrossRef CAS PubMed IUCr Journals Google Scholar
First citationSpackman, M. A., Byrom, P. G., Alfredsson, M. & Hermansson, K. (1999). Acta Cryst. A55, 30–47.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationStewart, R. F., Davidson, E. R. & Simpson, W. T. (1965). J. Chem. Phys. 42, 3175–3187.  CrossRef CAS Web of Science Google Scholar
First citationSun, Q., Zhang, X., Banerjee, S., Bao, P., Barbry, M., Blunt, N. S., Bogdanov, N. A., Booth, G. H., Chen, J., Cui, Z.-H., Eriksen, J. J., Gao, Y., Guo, S., Hermann, J., Hermes, M. R., Koh, K., Koval, P., Lehtola, S., Li, Z., Liu, J., Mardirossian, N., McClain, J. D., Motta, M., Mussard, B., Pham, H. Q., Pulkin, A., Purwanto, W., Robinson, P. J., Ronca, E., Sayfutyarova, E. R., Scheurer, M., Schurkus, H. F., Smith, J. E. T., Sun, C., Sun, S.-N., Upadhyay, S., Wagner, L. K., Wang, X., White, A., Whitfield, J. D., Williamson, M. J., Wouters, S., Yang, J., Yu, J. M., Zhu, T., Berkelbach, T. C., Sharma, S., Sokolov, A. Y. & Chan, G. K.-L. (2020). J. Chem. Phys. 153, 024109.  Web of Science CrossRef PubMed Google Scholar
First citationVolkov, A. M. P., Ferrugia, L. J., Gatti, C., Mallinson, P. R., Richter, T. & Koritsanszky, T. (2006). XD2006. University at Buffalo, State University of New York, NY, USA; Università di Milano, Italy; University of Glasgow, UK; CNR–ISTM, Milano, Italy; and Middle Tennessee State University, TN, USA.  Google Scholar
First citationWall, M. E. (2016). IUCrJ, 3, 237–246.  Web of Science CrossRef CAS PubMed IUCr Journals Google Scholar
First citationWieduwilt, E. K., Macetti, G., Malaspina, L. A., Jayatilaka, D., Grabowsky, S. & Genoni, A. (2020). J. Mol. Struct. 1209, 127934.  Web of Science CrossRef Google Scholar
First citationWoińska, M., Chodkiewicz, M. L. & Woźniak, K. (2021). Chem. Commun. 57, 3652–3655.  Google Scholar
First citationWoińska, M., Jayatilaka, D., Spackman, M. A., Edwards, A. J., Dominiak, P. M., Woźniak, K., Nishibori, E., Sugimoto, K. & Grabowsky, S. (2014). Acta Cryst. A70, 483–498.  Web of Science CSD CrossRef IUCr Journals Google Scholar

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

ISSN: 2052-2525