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

Journal logoSTRUCTURAL SCIENCE
CRYSTAL ENGINEERING
MATERIALS
ISSN: 2052-5206

TAAM: a reliable and user friendly tool for hydrogen-atom location using routine X-ray diffraction data

CROSSMARK_Color_square_no_text.svg

aBiological and Chemical Research Centre, Department of Chemistry, University of Warsaw, ul. Żwirki i Wigury 101, Warszawa, 02-089, Poland
*Correspondence e-mail: mchodkiewicz@chem.uw.edu.pl, pdomin@chem.uw.edu.pl

Edited by P. Macchi, University of Bern, Switzerland (Received 5 November 2019; accepted 2 March 2020; online 10 April 2020)

Hydrogen is present in almost all of the molecules in living things. It is very reactive and forms bonds with most of the elements, terminating their valences and enhancing their chemistry. X-ray diffraction is the most common method for structure determination. It depends on scattering of X-rays from electron density, which means the single electron of hydrogen is difficult to detect. Generally, neutron diffraction data are used to determine the accurate position of hydrogen atoms. However, the requirement for good quality single crystals, costly maintenance and the limited number of neutron diffraction facilities means that these kind of results are rarely available. Here it is shown that the use of Transferable Aspherical Atom Model (TAAM) instead of Independent Atom Model (IAM) in routine structure refinement with X-ray data is another possible solution which largely improves the precision and accuracy of X—H bond lengths and makes them comparable to averaged neutron bond lengths. TAAM, built from a pseudoatom databank, was used to determine the X—H bond lengths on 75 data sets for organic molecule crystals. TAAM parametrizations available in the modified University of Buffalo Databank (UBDB) of pseudoatoms applied through the DiSCaMB software library were used. The averaged bond lengths determined by TAAM refinements with X-ray diffraction data of atomic resolution (dmin ≤ 0.83 Å) showed very good agreement with neutron data, mostly within one single sample standard deviation, much like Hirshfeld atom refinement (HAR). Atomic displacements for both hydrogen and non-hydrogen atoms obtained from the refinements systematically differed from IAM results. Overall TAAM gave better fits to experimental data of standard resolution compared to IAM. The research was accompanied with development of software aimed at providing user-friendly tools to use aspherical atom models in refinement of organic molecules at speeds comparable to routine refinements based on spherical atom model.

1. Introduction

Hydrogen is the most abundant element in the universe. It is present in almost all molecules in living things (Emsley, 2011[Emsley, J. (2011). Nature's Building Blocks: an A-Z Guide to the Elements. Oxford University Press.]; Vladilo & Hassanali, 2018[Vladilo, G. & Hassanali, A. (2018). Life, 8, 1-22.]). It is highly reactive. It forms bonds with most elements, terminating their valences and enhancing the chemistry of those elements. Hydrogen bonding is responsible for the unique solvent properties of water, holding the two complementary strands of DNA together, the 3-D folding of proteins in enzymes and antibodies, binding of ligands to receptors and many other things. Hydrogen plays a significant role in many catalytic reactions both in biological systems and in small molecule chemistry (Taylor & Jacobsen, 2006[Taylor, M. S. & Jacobsen, E. N. (2006). Angew. Chem. Int. Ed. 45, 1520-1543.]; Doyle & Jacobsen, 2007[Doyle, A. G. & Jacobsen, E. N. (2007). Chem. Rev. 107, 5713-5743.]; Głowacki et al., 2013[Głowacki, E. D., Irimia-Vladu, M., Bauer, S. & Sariciftci, N. S. (2013). J. Mater. Chem. B, 1, 3742-3753.]; Belkova et al., 2016[Belkova, N. V., Epstein, L. M., Filippov, O. A. & Shubina, E. S. (2016). Chem. Rev. 116, 8545-8587.]). Understanding of the catalytic mechanisms of enzymes, the mechanism of protein folding and ligand–receptor complex formation demands accurate and precise determination of the position of the hydrogen atoms (Agback & Agback, 2018[Agback, P. & Agback, T. (2018). Sci. Rep. 8, 1-6.]; Chen et al., 2016[Chen, D., Oezguen, N., Urvil, P., Ferguson, C., Dann, S. M. & Savidge, T. C. (2016). Sci. Adv. 2, e1501240.]).

X-ray diffraction is the most commonly used method for chemical structure determination. This method depends on scattering of X-rays by the electron density, which is relatively low for hydrogen atoms; therefore, hydrogens are difficult to locate (Müller et al., 2006[Müller, P., Herbst-Irmer, R., Spek, A. L., Schneider, T. R. & Sawaya, M. R. (2006). Crystal Structure Refinement: A Crystallographer's Guide to SHELXL. Oxford University Press.]). To overcome this, neutron diffraction is used to determine the accurate position of hydrogen atoms, utilizing the fact that the neutron scattering length for hydrogen atoms is comparable to heavier elements. Most of the neutron diffraction experiments demand large single crystals (>1 mm3). Recent advancements in technology have made it possible to collect neutron data on smaller crystals (<1 mm3); however, the costly maintenance of neutron sources and the limited number of neutron diffraction facilities are significant bottlenecks for these kind of studies.

In general, routine diffraction data are modeled using an Independent Atom Model (IAM) which is based on individual atom contributions. The atoms in IAM are represented by spherically averaged electron density distributions obtained from quantum chemistry calculations for isolated atoms in their ground state (Coppens, 1997[Coppens, P. (1997). X-ray Charge Densities and Chemical Bonding. International Union of Crystallography and Oxford University Press.]). However, the electron density of an isolated hydrogen atom is significantly different from the one of a bonded hydrogen atom and the scattering factor obtained from IAM leads to meaningless bond lengths and other unrealistic properties (Stalke, 2012[Stalke, D. (2012). Electron Density and Chemical Bonding I: Experimental Charge Density Studies. Springer.]). The lack of core electrons in the hydrogen atom and polarization of its electron density towards the bonding regions (i.e. along the X—H bond, where X is any non-hydrogen atom), cause the X—H bond lengths determined for hydrogen atoms by IAM to be much shorter than expected. Additionally, the electron density in the vicinity of the proton in bonded hydrogen atoms is contracted compared to the isolated atom causing isotropic displacement parameters to be abnormally low (Stewart et al., 1965[Stewart, R. F., Davidson, E. R. & Simpson, W. T. (1965). J. Chem. Phys. 42, 3175-3187.]). Improvements to scattering factors for hydrogen, based on the ab initio electron density of the hydrogen molecule were proposed by Stewart, Davidson and Simpson (1965[Stewart, R. F., Davidson, E. R. & Simpson, W. T. (1965). J. Chem. Phys. 42, 3175-3187.]). The best spherical density model for bonded hydrogen recommended by them (SDS), along with IAM for other atoms, was used until now in most of the routine X-ray structure determination protocols (Wilson & Geist, 1993[Wilson, A. J. C. & Geist, V. (1993). Cryst. Res. Technol. 28, 110.]), which allow refinements to obtain more reliable isotropic displacement parameters for hydrogen atoms, but still lead to X—H bond lengths ≃20% shorter than the values obtained from neutron diffraction data (Stewart et al., 1965[Stewart, R. F., Davidson, E. R. & Simpson, W. T. (1965). J. Chem. Phys. 42, 3175-3187.]). Later it was noted that high-order refinement (dmin < 0.7 Å) which improves the accuracy of atomic positions (Ruysink & Vos, 1974[Ruysink, A. F. J. & Vos, A. (1974). Acta Cryst. A30, 503-506.]; Coppens, 1968[Coppens, P. (1968). Acta Cryst. B24, 1272-1274.]) also shows a considerable improvement in the X—H bond lengths (Hope & Ottersen, 1978[Hope, H. & Ottersen, T. (1978). Acta Cryst. B34, 3623-3626.]; Almlöf & Ottersen, 1979[Almlöf, J. & Ottersen, T. (1979). Acta Cryst. A35, 137-139.]). A further attempt by Stewart and co-workers was based on a fixed multipole expansion for a bonded hydrogen atom (i.e. polarized hydrogen atom) (Stewart et al., 1975[Stewart, R. F., Bentley, J. & Goodman, B. (1975). J. Chem. Phys. 63, 3786-3793.]). This model has largely improved the X—H bond lengths and has been successfully applied to a few systems (Roversi et al., 1996[Roversi, P., Merati, F., Destro, R. & Barzaghi, M. (1996). Can. J. Chem. 74, 1145-1161.]; Destro et al., 1988[Destro, R., Marsh, R. E. & Bianchi, R. (1988). J. Phys. Chem. 92, 966-973.], 2000[Destro, R., Roversi, P., Barzaghi, M. & Marsh, R. E. (2000). J. Phys. Chem. A, 104, 1047-1054.]; Madsen et al., 2004[Madsen, A. Ø., Sørensen, H. O., Flensburg, C., Stewart, R. F. & Larsen, S. (2004). Acta Cryst. A60, 550-561.]), but it was not yet as efficient or accurate to match the mean X—H bond lengths from neutron data. There are few other non-empirical methods available, such as extension of the X—H bond (Madsen et al., 2004[Madsen, A. Ø., Sørensen, H. O., Flensburg, C., Stewart, R. F. & Larsen, S. (2004). Acta Cryst. A60, 550-561.]; Overgaard et al., 2009[Overgaard, J., Platts, J. A. & Iversen, B. B. (2009). Acta Cryst. B65, 715-723.]), a combined approach, in which the high- and low-order refinement for non-hydrogen atoms and hydrogen atoms, respectively, along with the subsequent extension of the X—H bond is utilized to match the mean bond length values obtained from neutron data (Hoser et al., 2009[Hoser, A. A., Dominiak, P. M. & Woźniak, K. (2009). Acta Cryst. A65, 300-311.]).

It has been noted that the values of electron density parameters obtained from multipolar refinements based on Hansen and Coppens (1978[Hansen, N. K. & Coppens, P. (1978). Acta Cryst. A34, 909-921.]) formalism are almost identical for atoms in a similar chemical environment (Brock et al., 1991[Brock, C. P., Dunitz, J. D. & Hirshfeld, F. L. (1991). Acta Cryst. B47, 789-797.]). Therefore, a number of databanks were developed for different types of so-called transferable aspherical atoms (pseudoatoms) and the databanks were applied to create an electron density model called the Transferable Aspherical Atom Model (TAAM). The most common databanks are: ELMAM2 (Zarychta et al., 2007[Zarychta, B., Pichon-Pesme, V., Guillot, B., Lecomte, C. & Jelsch, C. (2007). Acta Cryst. A63, 108-125.]; Domagała et al., 2012[Domagała, S., Fournier, B., Liebschner, D., Guillot, B. & Jelsch, C. (2012). Acta Cryst. A68, 337-351.]; Nassour et al., 2017[Nassour, A., Domagala, S., Guillot, B., Leduc, T., Lecomte, C. & Jelsch, C. (2017). Acta Cryst. B73, 610-625.]), Invariom (Dittrich et al., 2004[Dittrich, B., Koritsánszky, T. & Luger, P. (2004). Angew. Chem. Int. Ed. 43, 2718-2721.], 2005[Dittrich, B., Hübschle, C. B., Messerschmidt, M., Kalinowski, R., Girnt, D. & Luger, P. (2005). Acta Cryst. A61, 314-320.], 2006[Dittrich, B., Hübschle, C. B., Luger, P. & Spackman, M. A. (2006). Acta Cryst. D62, 1325-1335.], 2013[Dittrich, B., Hübschle, C. B., Pröpper, K., Dietrich, F., Stolper, T. & Holstein, J. (2013). Acta Cryst. B69, 91-104.]) and UBDB (Volkov et al., 2007[Volkov, A., Messerschmidt, M. & Coppens, P. (2007). Acta Cryst. D63, 160-170.]; Dominiak et al., 2007[Dominiak, P. M., Volkov, A., Li, X., Messerschmidt, M. & Coppens, P. (2007). J. Chem. Theory Comput. 3, 232-247.]; Jarzembska & Dominiak, 2012[Jarzembska, K. N. & Dominiak, P. M. (2012). Acta Cryst. A68, 139-147.]; Kumar et al., 2019[Kumar, P., Gruza, B., Bojarowski, S. A. & Dominiak, P. M. (2019). Acta Cryst. A75, 398-408.]). In TAAM, instead of refining the parameters of atomic electron density, they are simply constrained to values typical for the corresponding atom type. The use of TAAM instead of IAM in the crystal structure refinement procedure with the standard diffraction [dmin ≤ 0.83 Å, sin θ/λ = 0.6 Å−1 recommended in the crystallographic journals (Spek, 2003[Spek, A. L. (2003). J. Appl. Cryst. 36, 7-13.], 2020[Spek, A. L. (2020). Acta Cryst. E76, 1-11.])] has largely improved the X—H bond lengths which became comparable to corresponding averaged neutron lengths (Dittrich et al., 2005[Dittrich, B., Hübschle, C. B., Messerschmidt, M., Kalinowski, R., Girnt, D. & Luger, P. (2005). Acta Cryst. A61, 314-320.], 2006[Dittrich, B., Munshi, P. & Spackman, M. A. (2006). Acta Cryst. C62, o633-o635.]; Bendeif & Jelsch, 2007[Bendeif, E. & Jelsch, C. (2007). Acta Cryst. C63, o361-o364.]; Domagała et al., 2011[Domagała, S., Munshi, P., Ahmed, M., Guillot, B. & Jelsch, C. (2011). Acta Cryst. B67, 63-78.]; Dadda et al., 2012[Dadda, N., Nassour, A., Guillot, B., Benali-Cherif, N. & Jelsch, C. (2012). Acta Cryst. A68, 452-463.]; Ahmed et al., 2011[Ahmed, M., Noureen, S., Gros, P. C., Guillot, B. & Jelsch, C. (2011). Acta Cryst. C67, o329-o333.]; Bąk et al., 2011[Bąk, J. M., Domagała, S., Hübschle, C., Jelsch, C., Dittrich, B. & Dominiak, P. M. (2011). Acta Cryst. A67, 141-153.]; Dittrich et al., 2017[Dittrich, B., Lübben, J., Mebs, S., Wagner, A., Luger, P. & Flaig, R. (2017). Chem. Eur. J. 23, 4605-4614.]). Other refinement parameters, such as reliability factor (Rf) and residual densities, are also improved (Dittrich et al., 2006[Dittrich, B., Munshi, P. & Spackman, M. A. (2006). Acta Cryst. C62, o633-o635.]; Volkov et al., 2004[Volkov, A., Li, X., Koritsanszky, T. & Coppens, P. (2004). J. Phys. Chem. A, 108, 4283-4300.]; Pichon-Pesme et al., 2004[Pichon-Pesme, V., Jelsch, C., Guillot, B. & Lecomte, C. (2004). Acta Cryst. A60, 204-208.]; Bąk et al., 2011[Bąk, J. M., Domagała, S., Hübschle, C., Jelsch, C., Dittrich, B. & Dominiak, P. M. (2011). Acta Cryst. A67, 141-153.]). However, these methods were limited to specific users and uses due to the complexity of their application. Despite the dedicated work of the software developers of MoPro (Domagała et al., 2012[Domagała, S., Fournier, B., Liebschner, D., Guillot, B. & Jelsch, C. (2012). Acta Cryst. A68, 337-351.]), Invariom Tools (Dittrich et al., 2004[Dittrich, B., Koritsánszky, T. & Luger, P. (2004). Angew. Chem. Int. Ed. 43, 2718-2721.], 2013[Dittrich, B., Hübschle, C. B., Pröpper, K., Dietrich, F., Stolper, T. & Holstein, J. (2013). Acta Cryst. B69, 91-104.]), LSDB (Volkov et al., 2004[Volkov, A., Li, X., Koritsanszky, T. & Coppens, P. (2004). J. Phys. Chem. A, 108, 4283-4300.]), XD (Volkov et al., 2006[Volkov, A., Macchi, P., Farrugia, L. J., Gatti, C., Mallinson, P., Richter, T. & Koritsanszky, T. (2006). XD2006. University at Buffalo, NY, USA, University of Milan, Italy, University of Glasgow, UK, CNRISTM, Milan, Italy, and Middle Tennessee State University, TN, USA.]), Jana2006 (Petříček et al., 2014[Petříček, V., Dušek, M. & Palatinus, L. (2014). Z. Kristallogr. Mater. 229, 345-352.]) and other highly specialized programs, TAAM refinement still required a high level of expertise to be reliably performed.

A new technique called Hirshfeld Atom Refinement (HAR), which is based on tailor-made aspherical atomic electron densities extracted from a crystal-field–embedded quantum-chemical electron density using Hirshfeld's scheme (Hirshfeld, 1977[Hirshfeld, F. L. (1977). Theor. Chim. Acta, 44, 129-138.]; Jayatilaka & Dittrich, 2008[Jayatilaka, D. & Dittrich, B. (2008). Acta Cryst. A64, 383-393.]; Capelli, Bürgi, Dittrich et al., 2014[Capelli, S. C., Bürgi, H.-B., Dittrich, B., Grabowsky, S. & Jayatilaka, D. (2014). IUCrJ, 1, 361-379.]; 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.]), was introduced. The general applicability of HAR in estimating the X—H bond lengths in small molecules was shown by Woinska et al. (2016[Woińska, M., Grabowsky, S., Dominiak, P. M., Woźniak, K. & Jayatilaka, D. (2016). Sci. Adv. 2, e1600192.]) where the accuracy and precision were comparable to mean averaged neutron bond lengths. The HAR method was implemented in the Tonto program (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 and A. Y. Zomaya, pp. 142-151. Berlin, Heidelberg: Springer.]) and later interfaced with Olex2 (Fugel et al., 2018[Fugel, M., Jayatilaka, D., Hupf, E., Overgaard, J., Hathwar, V. R., Macchi, P., Turner, M. J., Howard, J. A. K., Dolomanov, O. V., Puschmann, H., Iversen, B. B., Bürgi, H.-B. & Grabowsky, S. (2018). IUCrJ, 5, 32-44.]). Coupling HAR with the databases of extremely localized molecular orbitals [ELMO-DB; Meyer et al. (2016[Meyer, B., Guillot, B., Ruiz-Lopez, M. F. & Genoni, A. (2016). J. Chem. Theory Comput. 12, 1052-1067.]), Meyer & Genoni (2018[Meyer, B. & Genoni, A. (2018). J. Phys. Chem. A, 122, 8965-8981.])] reduced the time of computing the wavefunction for HAR without affecting the accuracy or the precision of the refinements (Malaspina et al., 2019[Malaspina, L. A., Wieduwilt, E. K., Bergmann, J., Kleemiss, F., Meyer, B., Ruiz-López, M. F., Pal, R., Hupf, E., Beckmann, J., Piltz, R. O., Edwards, A. J., Grabowsky, S. & Genoni, A. (2019). J. Phys. Chem. Lett. 10, 6973-6982.]). However, the computational cost of HARs and even of ELMO-DB HARs (when very large macromolecules are investigated) are still much higher compared to TAAM refinements. Another recent approach is the use of Bond Oriented Deformation Density (BODD) and Lone Pair Electron Density (LONE) via a tool called IDEAL (Lübben et al., 2019[Lübben, J., Wandtke, C. M., Hübschle, C. B., Ruf, M., Sheldrick, G. M. & Dittrich, B. (2019). Acta Cryst. A75, 50-62.]). This approach refines data as fast as IAM alone but does not reach the same sophistication in representing electron density as TAAM (including Invarioms) or HAR (Lübben et al., 2019[Lübben, J., Wandtke, C. M., Hübschle, C. B., Ruf, M., Sheldrick, G. M. & Dittrich, B. (2019). Acta Cryst. A75, 50-62.]).

While TAAM refinement is computationally more efficient, previous implementations of this method have not been used by a wide range of users who are not its developers. In this work, the applicability of TAAM in determining the X—H position with comparable accuracy and precision to neutron data was reinvestigated with the same 75 organic crystal data sets used by Woinska et al. (2016[Woińska, M., Grabowsky, S., Dominiak, P. M., Woźniak, K. & Jayatilaka, D. (2016). Sci. Adv. 2, e1600192.]). We focused our efforts on making TAAM-based refinements user friendly and also to provide a general software solution for determination of X—H bond lengths with high accuracy and precision for routine X-ray structure determination at dmin ≤ 0.83 Å. For this a new software implementation called DiSCaMB (Densities in Structural Chemistry and Molecular Biology) was developed to facilitate integration of the aspherical atom model into a wide range of refinement programs commonly used in X-ray crystallography for small and macromolecules (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.]). Here, the structures were modeled and refined using the TAAM-equipped library in DiSCaMB integrated with Olex2 (Dolomanov et al., 2009[Dolomanov, O. V., Bourhis, L. J., Gildea, R. J., Howard, J. A. K. & Puschmann, H. (2009). J. Appl. Cryst. 42, 339-341.]). The X—H bond lengths obtained from this were categorized and compared with the averaged neutron lengths (Allen & Bruno, 2010[Allen, F. H. & Bruno, I. J. (2010). Acta Cryst. B66, 380-386.]). A total of 24 types of X—H bonds were categorized for these 75 structures based on the hybridization and surrounding atoms of attached C, N, O, P atoms. Additionally O—H bonds in water molecules trapped in these structures were also included. Two sets of refinement were performed and related statistical comparison between the different models such as IAM, HAR (Woińska et al., 2016[Woińska, M., Grabowsky, S., Dominiak, P. M., Woźniak, K. & Jayatilaka, D. (2016). Sci. Adv. 2, e1600192.]) and TAAM were made, i.e at both the maximum resolution as obtained from the deposited structures and also at routine standard resolution of dmin = 0.8 Å. Additionally, anisotropic refinements of hydrogen-atom displacements were performed using TAAM.

2. Methodology

2.1. Implementation

The DiSCaMB library, capable of calculating scattering factors for Hansen–Coppens multipole model (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.]; Gildea et al., 2011[Gildea, R. J., Bourhis, L. J., Dolomanov, O. V., Grosse-Kunstleve, R. W., Puschmann, H., Adams, P. D. & Howard, J. A. K. (2011). J. Appl. Cryst. 44, 1259-1263.]), was extended to support TAAM by adding the capability to recognize atom types and handle TAAM parameters from a pseudoatom databank. A new format for a databank was developed and the UBDB2018 databank (Kumar et al., 2019[Kumar, P., Gruza, B., Bojarowski, S. A. & Dominiak, P. M. (2019). Acta Cryst. A75, 398-408.]) was translated to the new format. The details of the atom-typing algorithm implementation and the new databank format will be published elsewhere.

For the purpose of this study, a locally modified version of Olex2 (version 1.2) (Dolomanov et al., 2009[Dolomanov, O. V., Bourhis, L. J., Gildea, R. J., Howard, J. A. K. & Puschmann, H. (2009). J. Appl. Cryst. 42, 339-341.]) was used for the refinements. Among others, it incorporates the TAAM X-ray scattering factors available in the DiSCaMB library into the olex2.refine module.

The newly released version 1.3 of Olex2 permits the use of non-spherical atomic form factors provided via a text file (Midgley et al., 2019[Midgley, L., Bourhis, L. J., Dolomanov, O., Peyerimhoff, N. & Puschmann, H. (2019). arXiv: 1911.08847.]). Files for the kind of TAAM parameterization applied in this work can be generated with a program called discamb2tsc (available from https://4xeden.uw.edu.pl) and used for refinements with a publicly available version of Olex2 (Dolomanov et al., 2009[Dolomanov, O. V., Bourhis, L. J., Gildea, R. J., Howard, J. A. K. & Puschmann, H. (2009). J. Appl. Cryst. 42, 339-341.]). Results obtained with both implementations [(1) locally modified version of Olex2 version 1.2 and (2) discamb2tsc + Olex2 version 1.3] are essentially the same.

2.2. Crystal structure data sets

The starting geometries for 81 organic molecule crystal data sets were taken from the supplementary material of Woinska et al. (2016[Woińska, M., Grabowsky, S., Dominiak, P. M., Woźniak, K. & Jayatilaka, D. (2016). Sci. Adv. 2, e1600192.]), where the selection criteria for these structures are described. The reflection data sets were taken directly from original literature or Cambridge Structural Database (Groom et al., 2016[Groom, C. R., Bruno, I. J., Lightfoot, M. P. & Ward, S. C. (2016). Acta Cryst. B72, 171-179.]). In four cases (Nos. 9, 17, 43 and 77 in Table S1) the data were not publicly available. The reflection data were used for refinements as they were published, without any omission of reflections except for additional comparative analyses mentioned in Section 3.4[link] and for data set Nos 19, 27, 42 and 75. For the last four data sets, reflection omission was necessary to achieve refinement convergence.

2.3. IAM refinements

Structures were refined using the default IAM framework (Wilson & Geist, 1993[Wilson, A. J. C. & Geist, V. (1993). Cryst. Res. Technol. 28, 110.]) with olex2.refine from Olex2 (v1.2) using the diffraction data at maximum resolution and truncated at dmin = 0.8 Å. Hydrogen atoms were modeled with the SDS bonded hydrogen model and refined with isotropic displacement parameters, freely without any restraints or constraints. The averaged X—H bond lengths obtained were compared with the reported IAM results (Woińska et al., 2016[Woińska, M., Grabowsky, S., Dominiak, P. M., Woźniak, K. & Jayatilaka, D. (2016). Sci. Adv. 2, e1600192.]) obtained using SHELXL (Sheldrick, 2015[Sheldrick, G. M. (2015). Acta Cryst. C71, 3-8.]).

2.4. TAAM refinement

The structures obtained from IAM were further refined with olex2.refine integrated with TAAM. For the 13 missing atom types in the modified UBDB2018 databank (Kumar et al., 2019[Kumar, P., Gruza, B., Bojarowski, S. A. & Dominiak, P. M. (2019). Acta Cryst. A75, 398-408.]), the new atom types were defined and the corresponding aspherical parameters for atomic form factors were calculated. The procedure for addition of new atom types have been discussed elsewhere (Dominiak et al., 2007[Dominiak, P. M., Volkov, A., Li, X., Messerschmidt, M. & Coppens, P. (2007). J. Chem. Theory Comput. 3, 232-247.]; Kumar et al., 2019[Kumar, P., Gruza, B., Bojarowski, S. A. & Dominiak, P. M. (2019). Acta Cryst. A75, 398-408.]). In one case (No. 3 in Table S1) there were insufficient model molecules available in the CSD representing the three atom-type missing according to the currently used atom-typing algorithm in the UBDB and in another case (No. 72 in Table S1) the molecule contained an exceptionally strong intramolecular hydrogen bond which could not be accounted for within the current atom-typing algorithm (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.]). Out of the 81 structures, 75 are subjected to further analysis.

2.5. HAR data

The HAR data were taken as published in the supplementary material of Woinska et al. (2016[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.]).

2.6. Statistical analysis

All the statistical analyses were performed following the procedures described by Woinska et al. (2016[Woińska, M., Grabowsky, S., Dominiak, P. M., Woźniak, K. & Jayatilaka, D. (2016). Sci. Adv. 2, e1600192.]).

3. Results and discussions

3.1. Comparison of high-resolution X-ray IAM and TAAM refinements with neutron data for averaged X—H bond lengths

The initial IAM-based refinement comparison between the reported (Woińska et al., 2016[Woińska, M., Grabowsky, S., Dominiak, P. M., Woźniak, K. & Jayatilaka, D. (2016). Sci. Adv. 2, e1600192.]) and olex2.refine results showed that the X—H bond length were in very good agreement (Fig. S1), so there was no significant difference introduced by the refinement software. The statistical comparison of the different types of X—H bonds obtained from X-ray data after IAM, HAR (Woińska et al., 2016[Woińska, M., Grabowsky, S., Dominiak, P. M., Woźniak, K. & Jayatilaka, D. (2016). Sci. Adv. 2, e1600192.]) and TAAM refinements (with isotropic model of hydrogen-atom displacements) at 0.8 Å and also the maximum reported resolution were performed with neutron data as reference, Fig. 1[link]. Further statistical details are given in the SI (Table S1 and S2).

[Figure 1]
Figure 1
Comparison of the X—H bond lengths from X-ray and neutron data refinement at (a) maximum and (b) 0.8 Å resolution for the selected 75 structures. Hydrogen atoms were refined with isotropic displacement parameters for X-ray data. The X—H bond lengths obtained from the 75 structures were categorized, averaged and compared with the averaged neutron lengths as defined previously (Allen & Bruno, 2010[Allen, F. H. & Bruno, I. J. (2010). Acta Cryst. B66, 380-386.]). Additional O—H bonds in water molecules trapped in these structures were also included and the corresponding neutron lengths were taken from Woińska et. al. (2016[Woińska, M., Grabowsky, S., Dominiak, P. M., Woźniak, K. & Jayatilaka, D. (2016). Sci. Adv. 2, e1600192.]). HAR bond lengths for the 81 structures were taken from Woińska et al. (2016[Woińska, M., Grabowsky, S., Dominiak, P. M., Woźniak, K. & Jayatilaka, D. (2016). Sci. Adv. 2, e1600192.]). The layout of this figure follows the one proposed in Fig. 2[link] of the article by Woińska et al. (2016[Woińska, M., Grabowsky, S., Dominiak, P. M., Woźniak, K. & Jayatilaka, D. (2016). Sci. Adv. 2, e1600192.]). The bars in the figure show the SSDs of each associated bond type.

There is a significant X—H bond extension in TAAM refinement compared to IAM which caused the X—H bond lengths to approach the mean neutron values [Fig. 1[link](a)]. The differences between the mean bond lengths from the X-ray TAAM and neutron data were very low, typically one TAAM sample standard deviation (SSD) for all bond types. This error was reduced to less than one TAAM SSD for bond types with at least 30 representatives (Table S2). Mean bond lengths from IAM usually had a deviation of more than five IAM SSDs compared to neutron and TAAM lengths [Fig. 1[link](a)]. The difference in individual bond-type lengths from TAAM varies from 0.001 Å to 0.05 Å compared to neutron bond lengths. The averaged difference between TAAM and neutron mean lengths for all bond types was 0.025 Å which is comparable to the neutron SSDs (0.017 Å) averaged over all bond types. On the other hand, the mean IAM bond lengths were shorter by 0.12 Å on average compared to neutron lengths while the difference was an order of magnitude higher than for TAAM results. In the case of co-crystallized water molecules, which were usually disordered, the difference in the mean bond lengths was 0.04 Å with the SSD one third of the neutron SSD.

The precision of determining the X—H bond lengths can be further illustrated by examining the values of SSDs, as proposed by Woinska et al. (2016[Woińska, M., Grabowsky, S., Dominiak, P. M., Woźniak, K. & Jayatilaka, D. (2016). Sci. Adv. 2, e1600192.]). For NH2 (NH2 attached to sp2 carbon) and O—H bonds (aliphatic, aromatic alcohols and acids) TAAM SSDs were found to be approximately three times larger than neutron SSDs. However, in all these cases the number of observations was less than 30. In all other bond types, the TAAM SSDs were comparable to neutron SSDs (Figs. 1[link] and S2, Table S2).

While all these comparisons are based on the bond-type classes and may differ in specific individual structures due to the surrounding interactions involved, the disorder in the structures, or the data collection temperature, there was remarkably good agreement between the TAAM and neutron bond lengths.

3.2. Comparison of 0.8 Å resolution X-ray IAM and TAAM refinements with neutron data for averaged X—H bond lengths

The issues arising from data collection at high resolution such as lower intensity, merging, trailing of diffraction spots and other experimental errors cannot be ignored. With the aim of using TAAM for routine X-ray data refinement and to estimate accurate and precise X—H bond lengths, we performed TAAM refinement at dmin = 0.8 Å by truncating the high-resolution data and comparing the X—H bond lengths with the reference neutron lengths. This comparison is shown in Fig. 1[link](b). The trend remains the same with TAAM results showing the comparable accuracy and precision as at maximum resolution. The difference in individual mean bond lengths varies from 0.001 Å to 0.04 Å compared to neutron bond lengths. However, the differences in mean bond lengths between TAAM and neutron results was significantly reduced for some bond types (Table S2) compared to high-resolution results. This reduction was more prominent in the case of —OH and —NH bonds, irrespective of the number of representatives. For co-crystallized water molecules the difference in mean bond lengths also showed a greater than 50% reduction at 0.8 Å resolution compared to the maximum resolution (Table S2). This reduction in the differences of mean bond lengths at 0.8 Å resolution showed the further improvement in the accuracy of TAAM results with worsening resolution. While TAAM SSDs of most of the bond types improved with data truncation, there were slight increases for —OH and —NH bonds.

The trend of getting even more accurate results with aspherical scattering models when resolution of X-ray diffraction data is artificially truncated was already noted by others (Genoni et al., 2017[Genoni, A., Dos Santos, L. H. R., Meyer, B. & Macchi, P. (2017). IUCrJ, 4, 136-146.]). This has been explained by the fact that the information content about valence electron redistribution due to chemical bonding, intermolecular polarization and electron correlation is much higher in the low-order reflections than in high-order reflections. Resolution cut-offs change the balance between low-order and high-order reflections contributing to the minimized function, allowing the refinement procedure to achieve more accurate fitting for parameters that are heavily influenced by valence electrons.

3.3. Comparison of other refinement parameters for IAM and TAAM at 0.8 Å resolution

While modeling parameters such as the coordinates of hydrogen atoms from a fit to experimental data, it is important to validate the physical correctness of obtained parameters. Above we showed that TAAM produces more physically correct hydrogen atom coordinates than IAM, i.e. closer to neutron diffraction data. However, because TAAM was built from the UBDB, the generation of more correct hydrogen-atom coordinates from experimental X-ray data does not prove that TAAM is superior to other methods. UBDB was built from the electron densities of model organic molecules in which hydrogen atoms were positioned at the averaged neutron distances (Dominiak et al., 2007[Dominiak, P. M., Volkov, A., Li, X., Messerschmidt, M. & Coppens, P. (2007). J. Chem. Theory Comput. 3, 232-247.]). The fact that the X—H bond lengths from TAAM refinements were closer to neutron diffraction data may only be a result of the `what you put in is what you get out' effect. This is still acceptable as long as the proposed model not only leads to more physical parameters but also improved fits to experimental data. Here we investigate reliability factors (Rf) [here, R1 = ∑ ||Fo| − |Fc||/∑|Fo| with Fo > 4σ(Fo)] and residual Fourier electron density peaks to show if along with more accurate hydrogen-atom positions the improved fits from TAAM are achieved compared to IAM. In addition, the physical correctness of other fitted parameters such as atomic displacement parameters, should be analyzed. What is measured by an X-ray diffraction experiment is the static crystal electron density distorted by atomic displacements. Atomic displacement parameters may be modified during the least-squares fitting to compensate for weaknesses in the refined model (Gruza et al., 2020[Gruza, B., Chodkiewicz, M. L., Krzeszczakowska, J. & Dominiak, P. M. (2020). Acta Cryst. A76, 92-109.]). Achieving the proper deconvolution of atomic movements from the electron density is yet another indicator of quality for the electron density model.

The refinement of X-ray data at a standard 0.8 Å resolution with TAAM showed a one percent decrease in the Rf compared to IAM (Fig. 2[link]). However, on one occasion a large value of Rf was observed compared to other structures as seen in Fig. 2[link] with large blue and red spikes in both IAM and TAAM refinement. A closer look at the structure showed the presence of disorder (structure 19 with CH3 group disorder). The increase in Rf in the presence of disorder was more prominent in the case of TAAM refinement compared to other non-disordered structures, suggesting that the procedure was more efficient for disorder recognition and possible disorder refinement.

[Figure 2]
Figure 2
Comparison of reliability factor (Rf) for the TAAM and IAM refinements at 0.8 Å resolution. Hydrogen atoms were refined with isotropic displacement parameters.

A similar trend was observed in the case of residual (peaks and holes) Fourier electron densities obtained from IAM and TAAM. The residuals from TAAM showed an approximately 0.1 e Å−3 decrease compared to IAM as shown in Fig. 3[link]. In a few cases where the disorder was prominent, the residuals were found to be high in the case of TAAM refinement but lower than for IAM. These results further indicate the reliability of TAAM refinement and its sensitivity towards disorder compared to IAM.

[Figure 3]
Figure 3
Comparison of residual [i.e. peak (+) and hole (−)] electron densities (eÅ−3) for the TAAM and IAM refinement at 0.8 Å resolution. Hydrogen atoms were refined with isotropic displacement parameters. TAAM peak and hole are shown in blue while IAM peak and hole are shown in red.

We compared the atomic anisotropic displacement parameters (ADPs, here Uij) for non-hydrogen atoms which represent the averaged displacements of atoms from their mean positions in crystals. To find any general trends in the way ADPs from TAAM differ from those from IAM we performed an analysis of the overall size of the displacement ellipsoids by focusing on equivalent isotropic displacement parameters (Ueq) computed from Uij (Fischer & Tillmanns, 1988[Fischer, R. X. & Tillmanns, E. (1988). Acta Cryst. C44, 775-776.]). It appears that TAAM refinement leads to a systematic decrease in non-hydrogen Ueq compared to IAM (Fig. 4[link]). TAAM gives comparatively lower values and spread of Ueq with respect to IAM in non-disordered structures (Fig. 4[link]). In the case of disordered structures the spread of Ueq from TAAM refinement is slightly higher compared to other non-disordered structures but still lower than IAM. Over all structures, the averaged % difference between Ueq for non-hydrogen atoms from IAM and from TAAM is 13% (equation 1 in the supporting information).

[Figure 4]
Figure 4
Comparison of Ueq2) for non-hydrogen atoms obtained from TAAM and IAM refinements at 0.8 Å resolution. Hydrogen atoms were refined with isotropic displacement parameters.

Further analysis of Uiso for hydrogen atoms obtained from IAM and TAAM refinement showed systematically larger values for TAAM compared to IAM (Fig. 5[link]) which is opposite to what had previously been observed in the case of non-hydrogen atoms (Fig. 4[link]). The averaged % difference between Uiso of hydrogen atoms obtained from IAM and from TAAM was 32%.

[Figure 5]
Figure 5
Comparison of Uiso2) for hydrogen atoms obtained from TAAM and IAM refinements at 0.8 Å resolution. Hydrogen atoms were refined with isotropic displacement parameters.

Large scale validation of the Ueq/Uiso values with neutron data (i.e. analogous to the X—H bond length analysis) is much more difficult at the moment as analysis is limited by the amount of atomic displacement parameters available from neutron data. This information was not stored in the CSD and the quality of the available information strongly depends on quality of experimental data (Capelli, Bürgi, Dittrich et al., 2014[Capelli, S. C., Bürgi, H.-B., Dittrich, B., Grabowsky, S. & Jayatilaka, D. (2014). IUCrJ, 1, 361-379.]). Therefore here we decided to focus on selected data sets (23, 24, 49, 80, 81), for which analogous structures from neutron diffraction data that were measured at the same temperatures were available (Capelli, Bürgi, Dittrich et al., 2014[Capelli, S. C., Bürgi, H.-B., Dittrich, B., Grabowsky, S. & Jayatilaka, D. (2014). IUCrJ, 1, 361-379.]; Capelli, Bürgi, Mason & Jayatilaka, 2014[Capelli, S. C., Bürgi, H.-B., Mason, S. A. & Jayatilaka, D. (2014). Acta Cryst. C70, 949-952.]; Lübben et al., 2014[Lübben, J., Volkmann, C., Grabowsky, S., Edwards, A., Morgenroth, W., Fabbiani, F. P. A., Sheldrick, G. M. & Dittrich, B. (2014). Acta Cryst. A70, 309-316.]; Swaminathan et al., 1984[Swaminathan, S., Craven, B. M. & McMullan, R. K. (1984). Acta Cryst. B40, 300-306.]; Madsen et al., 2003[Madsen, A. Ø., Mason, S. & Larsen, S. (2003). Acta Cryst. B59, 653-663.]).

The results are not conclusive. For some structures (Nos. 80 and 81) Ueq obtained from TAAM for non-hydrogen atoms were found to be closer to neutron data compared to IAM, for others (Nos. 23, 24 and 49) the opposite is true [Fig. 6[link](a)]. Similar observations for HAR were reported by Capelli, Bürgi, Dittrich et al. (2014[Capelli, S. C., Bürgi, H.-B., Dittrich, B., Grabowsky, S. & Jayatilaka, D. (2014). IUCrJ, 1, 361-379.]). In the case of hydrogen atoms, the trend of Uiso from TAAM being systematically larger and closer to neutron data than IAM does not seem to depend on the data set and is observed for most of the hydrogen atoms [Fig. 6[link](b)]. The averaged % difference between Ueq/Uiso for hydrogen atoms obtained from IAM and from neutron data was 50% while the averaged % difference between TAAM and neutron data was only 0.5%.

[Figure 6]
Figure 6
Comparison of Ueq/Uiso2) for (a) non-hydrogen and (b) hydrogen atoms obtained from reported neutron data and IAM and TAAM refinements with X-ray data at 0.8 Å resolution for selected structures (23, 24, 49, 80, 81). Hydrogen atoms were refined with isotropic displacement parameters.

As reported earlier, ADPs from a conventional spherical atom refinement (IAM) are systematically contaminated by bonding and lone-pair electron densities (Dittrich et al., 2008[Dittrich, B., McKinnon, J. J. & Warren, J. E. (2008). Acta Cryst. B64, 750-759.]; Bąk et al., 2011[Bąk, J. M., Domagała, S., Hübschle, C., Jelsch, C., Dittrich, B. & Dominiak, P. M. (2011). Acta Cryst. A67, 141-153.]; Domagała et al., 2011[Domagała, S., Munshi, P., Ahmed, M., Guillot, B. & Jelsch, C. (2011). Acta Cryst. B67, 63-78.]). However, validations of ADPs was often compromised by the quality of experimental data (Dittrich et al., 2008[Dittrich, B., McKinnon, J. J. & Warren, J. E. (2008). Acta Cryst. B64, 750-759.]; Capelli, Bürgi, Dittrich et al., 2014[Capelli, S. C., Bürgi, H.-B., Dittrich, B., Grabowsky, S. & Jayatilaka, D. (2014). IUCrJ, 1, 361-379.]). Analyses performed not only for high-quality high-resolution experimental X-ray and neutron data but also for simulated X-ray data shed more light on the topic (Gruza et al., 2020[Gruza, B., Chodkiewicz, M. L., Krzeszczakowska, J. & Dominiak, P. M. (2020). Acta Cryst. A76, 92-109.]). For the simulated data the targeted values of Ueq/Uiso to be achieved during refinements were predetermined. The Ueq/Uiso values from TAAM refinements at 0.8 Å were much closer to the target values compared to IAM. Nevertheless, large-scale studies are needed.

3.4. Comparison of TAAM and HAR results

It is interesting to compare TAAM results to HAR. Before we discuss the results of that comparison, it must be noted that the observed differences do not rely only upon the differences in atomic scattering factors computed either from TAAM or from Hirshfeld atoms. In most of the TAAM refinements we used the reflection data as they were originally published, whereas for HAR, thousands of reflections were omitted from the refinement to achieve convergence. While the general consensus is to use all of the collected intensities in refinement, omission of the weakest intensities sometimes improves the refinement depending upon the method used (i.e refinements against |F| or F2) (Merli, 2002[Merli, M. (2002). Z. Kristallogr. 217, 103-108.]). Moreover, there are many more differences between our TAAM refinements and published HAR; data were refined against structure factor magnitudes |F| (HAR) or F2 (TAAM) and the weighting schemes were different. Only when the same refinement strategy with the same software implementation is used, the effects of replacing TAAM scattering factors by Hirshfeld atoms can been investigated more deeply.

The values obtained for the averaged X—H bond lengths from TAAM are within the same magnitude range as those of HAR bond lengths, though HAR results are systematically slightly closer to neutron data than TAAM (Fig. 1[link]). For refinements at 0.8 Å resolution with isotropic hydrogen displacement parameters, the mean TAAM bond lengths averaged over all bond types were shorter by 0.020 Å compared to neutron lengths whereas the averaged shortening for HAR was 0.014 Å. The importance of the difference might be further judged from the SSDs of TAAM and HAR results, which were compared by plotting the difference between the TAAM and HAR mean bond lengths along with HAR and TAAM SSDs (Fig. S3). The SSDs from TAAM are comparable to HAR in most cases and in the remaining cases they were either higher (mostly bond types belonging to Z—O—H and Z—Csp3—H3 groups) or lower than the HAR SSDs. Only on a few occasions was the mean X—H bond length difference between TAAM and HAR larger than one SSD.

The refinement indicators at standard 0.8 Å resolution such as Rf and residual densities from TAAM were also comparable to HAR (Figs. S4 and S5). In most cases the differences in HAR and TAAM indicators were much smaller than the differences between any of these two methods and IAM, although the HAR results usually were slightly better. The Ueq parameters for non-hydrogen atoms from HAR were comparable to TAAM, the averaged percentage difference between TAAM and HAR Ueq was found to be 0.8% (Fig. S6). For hydrogen atoms the averaged percentage difference between TAAM and HAR Uiso was found to be 7% (Fig. S7). Further studies are needed to investigate which method, HAR or TAAM, gives Ueq/Uiso closer to reality.

Even more detailed comparison between HAR and TAAM refinement was performed on four structures (27, 36, 42, 75) where the reflections were removed as was done in the case of HAR. Here results for both refinement resolutions (maximum or 0.8 Å) as well as both types of hydrogen-atom displacements (isotropic or anisotropic) were analyzed. In all cases, the refinement parameters (Rf, residual densities) and Ueq for non-hydrogen atoms thus obtained from TAAM were very similar to HAR (Table S3, Fig. S8).

3.5. Anisotropic refinement of hydrogen-atom displacements using TAAM approach

It was long believed that it was not possible to obtain reliable anisotropic displacement parameters for hydrogen atoms directly from standard X-ray diffraction data. There are various tools available for estimation of hydrogen ADPs. These are, for example, the segmented-body translation libration screw (TLS) refinement method along with Invariom approach used in the ADP-toolkit program (Lübben et al., 2015[Lübben, J., Bourhis, L. J. & Dittrich, B. (2015). J. Appl. Cryst. 48, 1785-1793.]), or the SHADE method using ADPs from neutron experiments to approximate hydrogen ADPs and available from the SHADE web server (Madsen et al., 2003[Madsen, A. Ø., Mason, S. & Larsen, S. (2003). Acta Cryst. B59, 653-663.]; Madsen & Hoser, 2014[Madsen, A. Ø. & Hoser, A. A. (2014). J. Appl. Cryst. 47, 2100-2104.]). It was also shown that for sub-atomic resolution X-ray data of exceptional quality, the multipolar model with polarized hydrogen atoms and additional bond-directed dipoles also lead to hydrogen ADPs close to the neutron values (Zhurov et al., 2011[Zhurov, V. V., Zhurova, E. A., Stash, A. I. & Pinkerton, A. A. (2011). Acta Cryst. A67, 160-173.]). Refinements of hydrogen ADPs with X-ray data using the HAR method livened-up the discussion, as the authors proposed that HAR may give reliable hydrogen ADPs (Jayatilaka & Dittrich, 2008[Jayatilaka, D. & Dittrich, B. (2008). Acta Cryst. A64, 383-393.]; Capelli, Bürgi, Dittrich et al., 2014[Capelli, S. C., Bürgi, H.-B., Dittrich, B., Grabowsky, S. & Jayatilaka, D. (2014). IUCrJ, 1, 361-379.]). A recent comparison of different strategies in modeling the hydrogen atoms indicates that although HAR yields reliable C—H bond lengths, caution should be taken if hydrogen ADPs are to be obtained with this method (Köhler et al., 2019[Köhler, C., Lübben, J., Krause, L., Hoffmann, C., Herbst-Irmer, R. & Stalke, D. (2019). Acta Cryst. B75, 434-441.]).

To provide more results to the discussion we decided to check if it is possible to routinely obtain hydrogen ADPs from TAAM refinements with X-ray data and how obtained parameters compare with other methods. Anisotropic refinements of hydrogen atoms were performed on 75 structures using TAAM. In most cases hydrogen ADPs were successfully refined. Successful refinements were those that were stable and convergence was achieved, and non-positive definite (NPD) ADPs were not observed or observed for only one or two atoms per structure. Statistical analyses analogous to the one above for isotropic refinements were also performed. The averaged X—H bond lengths were comparable to averaged neutron bond lengths (Fig. S9); however, noticeable changes were visible compared to isotropic hydrogen refinement. The averaged X—H bond lengths improved in some cases but not others (Fig. S10), which was consistent with a similar observation for HAR (Woińska et al., 2016[Woińska, M., Grabowsky, S., Dominiak, P. M., Woźniak, K. & Jayatilaka, D. (2016). Sci. Adv. 2, e1600192.]). There were marginal to significant improvements in Rf and residuals in anisotropic hydrogen refinement compared to isotropic hydrogen refinement at 0.8 Å resolution (Figs. S11 and S12). The refinement parameters obtained from TAAM were also comparable to the HAR results at 0.8 Å resolution (Figs. S13 and S14). The hydrogen ADPs were obtained from information derived from the X-ray data for a particular compound and hence the quality of the ADPs is poor in some cases, more often with obliqueness and oblateness. Out of all the anisotropically refined hydrogen atoms (949) only 4% (41) were found to be showing NPD at 0.8 Å resolution and 1% (10) at maximum reported resolution, which is consistent with the HAR results (Woińska et al., 2016[Woińska, M., Grabowsky, S., Dominiak, P. M., Woźniak, K. & Jayatilaka, D. (2016). Sci. Adv. 2, e1600192.]). The NPDs were found mostly in those cases where the water molecules were disordered or the data quality was poor.

In order to further investigate and verify the correctness of the hydrogen ADPs, comparisons were drawn between the hydrogen Uij values obtained from neutron, HAR and TAAM refinements, respectively, at 0.8 Å resolution, made for five representative crystal structures, Nos. 23, 24, 49, 80 and 81. A comparison for Uij values shows a good correlation (R2 = 0.70) between TAAM and neutron Uij values which was comparable to the correlation between neutron and HAR (R2 = 0.78) (Fig. S15). The correlation coefficient between TAAM and HAR Uij was found to be 0.87. A negligible change in the correlation coefficient was observed when the data with the omitted reflections were used (Fig. S15). Further we compared the ADPs (Uij) for all the atoms from structures 23, 24, 49, 80 and 81 using similarity index (S12) (Whitten & Spackman, 2006[Whitten, A. E. & Spackman, M. A. (2006). Acta Cryst. B62, 875-888.]). For two ADP tensors, the lower the S12 value was, the higher the observed similarity was, S12 = 0 means that these two tensors were exactly the same. We observed (Table S4) that TAAM ADPs for non-hydrogen atoms showed similarity (S12) with neutron ADPs comparable to the HAR results reported by Woińska et al. (2016[Woińska, M., Grabowsky, S., Dominiak, P. M., Woźniak, K. & Jayatilaka, D. (2016). Sci. Adv. 2, e1600192.]), and this was true for all five structures and both versions of refinements (with or without omissions of reflections in TAAM). For hydrogen ADPs, ignoring the NPDs, similarity indexes obtained for TAAM and HAR were also somewhat comparable.

4. Conclusions

In conclusion, we have confirmed the high accuracy and precision of mean X—H bond lengths obtained from TAAM, which were comparable to the mean neutron bond lengths, both at the maximum reported and 0.8 Å resolution. At 0.8 Å resolution the accuracy further improved compared to high resolution, especially in the case of the polar hydrogen atoms in NH and OH bond types. The TAAM refinement showed sensitivity towards disorder with high values of refinement parameter indicators. A comparison at 0.8 Å resolution between the IAM and TAAM refinement showed a significant improvement in refinement parameter indicators such as Rf values and residual electron densities for TAAM. A systematic and significant reduction in displacement parameters for non-hydrogen atoms and enlargement for hydrogen atoms was observed in TAAM refinements at 0.8 Å compared to IAM. Limited comparisons to neutron data suggested that atomic displacements for hydrogen atoms from TAAM were physically more meaningful than from IAM; further studies are needed to validate non-hydrogen-atom displacement parameters. The comparison of refinement parameter indicators for TAAM and HAR showed good agreement. All the results showed that the accuracy and precision of X—H bond lengths from TAAM refinement were not much worse than those reported from HAR and the obtained bond lengths were not worse than the reference neutron bond distances. It was indeed possible to refine hydrogen displacement parameters anisotropically using TAAM with good quality X-ray data and the results obtained were satisfactorily close to neutron values. Introduction of anisotropic hydrogen atoms did not appear to improve the other refinement parameters; further studies are need to confirm that the procedure does not cause overfitting.

These encouraging results establish the usefulness of TAAM for routine structure determination using X-ray data which is accessible to scientific users interested in the study of hydrogen bonds. TAAM refinements localize hydrogen atoms almost as accurately and precisely as neutron diffraction, overcoming deficiencies of IAM, as showed by bond lengths and displacement parameters. Moreover, the integration of TAAM methodology with Olex2, which is available for structure refinement to the crystallography community, makes TAAM a reliable and user-friendly tool to be applied in the last steps of refinement during routine X-ray structure determination of organic molecule crystals. Application of TAAM to polymer network structures, disordered compounds (Bąk et al., 2009[Bąk, J. M., Dominiak, P. M., Wilson, C. C. & Woźniak, K. (2009). Acta Cryst. A65, 490-500.]; Dittrich et al., 2016[Dittrich, B., Schürmann, C. & Hübschle, C. B. (2016). Z. Kristallogr. 231, 725-736.]), organic fragments in organometallic complexes (Wandtke et al., 2017[Wandtke, C. M., Weil, M., Simpson, J. & Dittrich, B. (2017). Acta Cryst. B73, 794-804.]) or molecular biology (Jelsch et al., 1998[Jelsch, C., Pichon-Pesme, V., Lecomte, C. & Aubry, A. (1998). Acta Cryst. D54, 1306-1318.]; Malinska & Dauter, 2016[Malinska, M. & Dauter, Z. (2016). Acta Cryst. D72, 770-779.]) seems to be possible and it will be investigated in future works. General applicability of TAAM from a databank for metal atoms and inorganic structures is currently not possible with the existing atom-typing algorithm and multipolar model parametrization used to build the databank, and requires further feasibility studies. Another missing feature of the current TAAM implementation is a lack of crystal field effects in the model built from UBDB, which in the future may possibly be overcome by implementation of the ELMAM2 database.

5. Related literature

The following references are cited in the supporting information: Munshi et al. (2006[Munshi, P., Thakur, T. S., Guru Row, T. N. & Desiraju, G. R. (2006). Acta Cryst. B62, 118-127.]); Zarychta et al. (2011[Zarychta, B., Zaleski, J., Kyzioł, J., Daszkiewicz, Z. & Jelsch, C. (2011). Acta Cryst. B67, 250-262.]); Chen et al. (2007[Chen, Y.-S., Stash, A. I. & Pinkerton, A. A. (2007). Acta Cryst. B63, 309-318.]); Nassour et al. (2014[Nassour, A., Kubicki, M., Wright, J., Borowiak, T., Dutkiewicz, G., Lecomte, C. & Jelsch, C. (2014). Acta Cryst. B70, 197-211.]); Slouf et al. (2002[Slouf, M., Holy, A., Petříček, V. & Cisarova, I. (2002). Acta Cryst. B58, 519-529.]); Munshi & Guru Row (2002[Munshi, P. & Guru Row, T. N. (2002). Acta Cryst. B58, 1011-1017.], 2006[Munshi, P. & Guru Row, T. N. (2006). Acta Cryst. B62, 612-626.]); Bianchi et al. (1998[Bianchi, R., Gervasio, G. & Viscardi, G. (1998). Acta Cryst. B54, 66-72.]); Janicki & Starynowicz (2010)[Janicki, R. & Starynowicz, P. (2010). Acta Cryst. B66, 559-567.]; Bürgi et al. (2002[Bürgi, H.-B., Capelli, S. C., Goeta, A. E., Howard, J. A. K., Spackman, M. A. & Yufit, D. S. (2002). Chem. Eur. J. 8, 3512-3521.]); Zhurova & Pinkerton (2001[Zhurova, E. A. & Pinkerton, A. A. (2001). Acta Cryst. B57, 359-365.]); Małecka et al. (2010[Małecka, M., Chęcińska, L., Rybarczyk-Pirek, A., Morgenroth, W. & Paulmann, C. (2010). Acta Cryst. B66, 687-695.]); Dittrich et al. (2009[Dittrich, B., Hübschle, C. B., Holstein, J. J. & Fabbiani, F. P. A. (2009). J. Appl. Cryst. 42, 1110-1121.], 2002[Dittrich, B., Koritsánszky, T., Grosche, M., Scherer, W., Flaig, R., Wagner, A., Krane, H. G., Kessler, H., Riemer, C., Schreurs, A. M. M. & Luger, P. (2002). Acta Cryst. B58, 721-727.], 2012[Dittrich, B., Sze, E., Holstein, J. J., Hübschle, C. B. & Jayatilaka, D. (2012). Acta Cryst. A68, 435-442.], 2007[Dittrich, B., Munshi, P. & Spackman, M. A. (2007). Acta Cryst. B63, 505-509.]); Meents et al. (2008[Meents, A., Dittrich, B., Johnas, S. K. J., Thome, V. & Weckert, E. F. (2008). Acta Cryst. B64, 42-49.]); Bibila Mayaya Bisseyou et al. (2012[Bibila Mayaya Bisseyou, Y., Bouhmaida, N., Guillot, B., Lecomte, C., Lugan, N., Ghermani, N. & Jelsch, C. (2012). Acta Cryst. B68, 646-660.]); Gulliot et al. (2001[Guillot, R., Muzet, N., Dahaoui, S., Lecomte, C. & Jelsch, C. (2001). Acta Cryst. B57, 567-578.]); Flaig et al. (1998[Flaig, R., Koritsanszky, T., Zobel, D. & Luger, P. (1998). J. Am. Chem. Soc. 120, 2227-2238.], 1999[Flaig, R., Koritsánszky, T., Janczak, J., Krane, H.-G., Morgenroth, W. & Luger, P. (1999). Angew. Chem. Int. Ed. 38, 1397-1400.]); Tsirelson et al. (2006[Tsirelson, V. G., Stash, A. I., Potemkin, V. A., Rykounov, A. A., Shutalev, A. D., Zhurova, E. A., Zhurov, V. V., Pinkerton, A. A., Gurskaya, G. V. & Zavodnik, V. E. (2006). Acta Cryst. B62, 676-688.]); Farrugia et al. (2002[Farrugia, L. J., Kočovský, P., Senn, H. M. & Vyskočil, S. (2009). Acta Cryst. B65, 757-769.]); Pichon-Pesme et al. (2000[Pichon-Pesme, V., Lachekar, H., Souhassou, M. & Lecomte, C. (2000). Acta Cryst. B56, 728-737.]); Benabicha et al. (2000[Benabicha, F., Pichon-Pesme, V., Jelsch, C., Lecomte, C. & Khmou, A. (2000). Acta Cryst. B56, 155-165.]); Śled et al. (2010[Śledź, P., Kamiński, R., Chruszcz, M., Zimmerman, M. D., Minor, W. & Woźniak, K. (2010). Acta Cryst. B66, 482-492.]); Chęcińska et al. (2011[Chęcińska, L., Grabowsky, S., Małecka, M., Rybarczyk-Pirek, A. J., Jóźwiak, A., Paulmann, C. & Luger, P. (2011). Acta Cryst. B67, 569-581.], 2006[Chęcińska, L., Förster, D., Morgenroth, W. & Luger, P. (2006). Acta Cryst. C62, o454-o457.]); Förster et al. (2005[Förster, D., Messerschmidt, M. & Luger, P. L. (2005). Acta Cryst. C61, o420-o421.]); Grabowsky et al. (2005[Grabowsky, S., Kalinowski, R., Weber, M., Förster, D., Paulmann, C. & Luger, P. (2009). Acta Cryst. B65, 488-501.]); Kalinowski et al. (2007[Kalinowski, R., Dittrich, B., Hübschle, C. B., Paulmann, C. & Luger, P. (2007). Acta Cryst. B63, 753-767.]); Scheins et al. (2004[Scheins, S., Dittrich, B., Messerschmidt, M., Paulmann, C. & Luger, P. (2004). Acta Cryst. B60, 184-190.], 2005[Scheins, S., Messerschmidt, M. & Luger, P. (2005). Acta Cryst. B61, 443-448.]); Lübben et al. (2014[Lübben, J., Volkmann, C., Grabowsky, S., Edwards, A., Morgenroth, W., Fabbiani, F. P. A., Sheldrick, G. M. & Dittrich, B. (2014). Acta Cryst. A70, 309-316.]); Dahaoui et al. (1999[Dahaoui, S., Jelsch, C., Howard, J. A. K. & Lecomte, C. (1999). Acta Cryst. B55, 226-230.]); Kamiński et al. (2014[Kamiński, R., Domagała, S., Jarzembska, K. N., Hoser, A. A., Sanjuan-Szklarz, W. F., Gutmann, M. J., Makal, A., Malińska, M., Bąk, J. M. & Woźniak, K. (2014). Acta Cryst. A70, 72-91.]); Paul et al. (2011[Paul, A., Kubicki, M., Jelsch, C., Durand, P. & Lecomte, C. (2011). Acta Cryst. B67, 365-378.]); Poulain et al. (2014[Poulain, A., Wenger, E., Durand, P., Jarzembska, K. N., Kamiński, R., Fertey, P., Kubicki, M. & Lecomte, C. (2014). IUCrJ, 1, 110-118.]); Zhurova et al. (2006[Zhurova, E. A., Tsirelson, V. G., Zhurov, V. V., Stash, A. I. & Pinkerton, A. A. (2006). Acta Cryst. B62, 513-520.]); Holstein et al. (2010[Holstein, J. J., Luger, P., Kalinowski, R., Mebs, S., Paulman, C. & Dittrich, B. (2010). Acta Cryst. B66, 568-577.]); Dittrich & Spackman (2007[Dittrich, B. & Spackman, M. A. (2007). Acta Cryst. A63, 426-436.]); Jaradat et al. (2007[Jaradat, D. M. M., Mebs, S., Chęcińska, L. & Luger, P. (2007). Carbohydr. Res. 342, 1480-1489.]); Hübschle et al. (2008[Hübschle, C. B., Dittrich, B., Grabowsky, S., Messerschmidt, M. & Luger, P. (2008). Acta Cryst. B64, 363-374.]); Birkedal et al. (2004[Birkedal, H., Madsen, D., Mathiesen, R. H., Knudsen, K., Weber, H.-P., Pattison, P. & Schwarzenbach, D. (2004). Acta Cryst. A60, 371-381.]).

Supporting information


Acknowledgements

MLC thanks Luc Bourhis, Oleg Dolomanov and Horst Puschmann for information/comments on integration of DiSCaMB functionality into Olex2. KKJ thanks Magdalena Woińska for disclosing all the methodological details regarding the work presented by Woińska et al. (2016[Woińska, M., Grabowsky, S., Dominiak, P. M., Woźniak, K. & Jayatilaka, D. (2016). Sci. Adv. 2, e1600192.]).

Funding information

The following funding is acknowledged: the National Centre of Science (Poland) (grant OPUS No. UMO-2017/27/B/ST4/02721 to Paulina Maria Dominiak) and PL-Grid Infrastructure through grant UBDB2019.

References

First citationAgback, P. & Agback, T. (2018). Sci. Rep. 8, 1–6.  Web of Science CrossRef CAS PubMed Google Scholar
First citationAhmed, M., Noureen, S., Gros, P. C., Guillot, B. & Jelsch, C. (2011). Acta Cryst. C67, o329–o333.  Web of Science CSD CrossRef IUCr Journals Google Scholar
First citationAllen, F. H. & Bruno, I. J. (2010). Acta Cryst. B66, 380–386.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationAlmlöf, J. & Ottersen, T. (1979). Acta Cryst. A35, 137–139.  CrossRef IUCr Journals Web of Science Google Scholar
First citationBąk, J. M., Domagała, S., Hübschle, C., Jelsch, C., Dittrich, B. & Dominiak, P. M. (2011). Acta Cryst. A67, 141–153.  Web of Science CrossRef IUCr Journals Google Scholar
First citationBąk, J. M., Dominiak, P. M., Wilson, C. C. & Woźniak, K. (2009). Acta Cryst. A65, 490–500.  Web of Science CSD CrossRef IUCr Journals Google Scholar
First citationBelkova, N. V., Epstein, L. M., Filippov, O. A. & Shubina, E. S. (2016). Chem. Rev. 116, 8545–8587.  Web of Science CrossRef CAS PubMed Google Scholar
First citationBenabicha, F., Pichon-Pesme, V., Jelsch, C., Lecomte, C. & Khmou, A. (2000). Acta Cryst. B56, 155–165.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationBendeif, E. & Jelsch, C. (2007). Acta Cryst. C63, o361–o364.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationBianchi, R., Gervasio, G. & Viscardi, G. (1998). Acta Cryst. B54, 66–72.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationBibila Mayaya Bisseyou, Y., Bouhmaida, N., Guillot, B., Lecomte, C., Lugan, N., Ghermani, N. & Jelsch, C. (2012). Acta Cryst. B68, 646–660.  Web of Science CSD CrossRef IUCr Journals Google Scholar
First citationBirkedal, H., Madsen, D., Mathiesen, R. H., Knudsen, K., Weber, H.-P., Pattison, P. & Schwarzenbach, D. (2004). Acta Cryst. A60, 371–381.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationBrock, C. P., Dunitz, J. D. & Hirshfeld, F. L. (1991). Acta Cryst. B47, 789–797.  CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationBürgi, H.-B., Capelli, S. C., Goeta, A. E., Howard, J. A. K., Spackman, M. A. & Yufit, D. S. (2002). Chem. Eur. J. 8, 3512–3521.  CrossRef PubMed CAS 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 citationCapelli, S. C., Bürgi, H.-B., Mason, S. A. & Jayatilaka, D. (2014). Acta Cryst. C70, 949–952.  Web of Science CSD CrossRef IUCr Journals Google Scholar
First citationChęcińska, L., Förster, D., Morgenroth, W. & Luger, P. (2006). Acta Cryst. C62, o454–o457.  Web of Science CSD CrossRef IUCr Journals Google Scholar
First citationChęcińska, L., Grabowsky, S., Małecka, M., Rybarczyk-Pirek, A. J., Jóźwiak, A., Paulmann, C. & Luger, P. (2011). Acta Cryst. B67, 569–581.  Web of Science CSD CrossRef IUCr Journals Google Scholar
First citationChen, D., Oezguen, N., Urvil, P., Ferguson, C., Dann, S. M. & Savidge, T. C. (2016). Sci. Adv. 2, e1501240.  Web of Science CrossRef PubMed Google Scholar
First citationChen, Y.-S., Stash, A. I. & Pinkerton, A. A. (2007). Acta Cryst. B63, 309–318.  Web of Science CSD CrossRef 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 citationCoppens, P. (1968). Acta Cryst. B24, 1272–1274.  CrossRef CAS IUCr Journals Web of Science Google Scholar
First citationCoppens, P. (1997). X-ray Charge Densities and Chemical Bonding. International Union of Crystallography and Oxford University Press.  Google Scholar
First citationDadda, N., Nassour, A., Guillot, B., Benali-Cherif, N. & Jelsch, C. (2012). Acta Cryst. A68, 452–463.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationDahaoui, S., Jelsch, C., Howard, J. A. K. & Lecomte, C. (1999). Acta Cryst. B55, 226–230.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationDestro, R., Marsh, R. E. & Bianchi, R. (1988). J. Phys. Chem. 92, 966–973.  CSD CrossRef CAS Web of Science Google Scholar
First citationDestro, R., Roversi, P., Barzaghi, M. & Marsh, R. E. (2000). J. Phys. Chem. A, 104, 1047–1054.  Web of Science CSD CrossRef CAS Google Scholar
First citationDittrich, B., Hübschle, C. B., Holstein, J. J. & Fabbiani, F. P. A. (2009). J. Appl. Cryst. 42, 1110–1121.  Web of Science CSD CrossRef CAS IUCr Journals 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., Hübschle, C. B., Pröpper, K., Dietrich, F., Stolper, T. & Holstein, J. (2013). Acta Cryst. B69, 91–104.  CrossRef IUCr Journals Google Scholar
First citationDittrich, B., Koritsánszky, T., Grosche, M., Scherer, W., Flaig, R., Wagner, A., Krane, H. G., Kessler, H., Riemer, C., Schreurs, A. M. M. & Luger, P. (2002). Acta Cryst. B58, 721–727.  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., Lübben, J., Mebs, S., Wagner, A., Luger, P. & Flaig, R. (2017). Chem. Eur. J. 23, 4605–4614.  Web of Science CrossRef CAS PubMed Google Scholar
First citationDittrich, B., McKinnon, J. J. & Warren, J. E. (2008). Acta Cryst. B64, 750–759.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationDittrich, B., Munshi, P. & Spackman, M. A. (2006). Acta Cryst. C62, o633–o635.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationDittrich, B., Munshi, P. & Spackman, M. A. (2007). Acta Cryst. B63, 505–509.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationDittrich, B., Schürmann, C. & Hübschle, C. B. (2016). Z. Kristallogr. 231, 725–736.  CAS Google Scholar
First citationDittrich, B. & Spackman, M. A. (2007). Acta Cryst. A63, 426–436.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationDittrich, B., Sze, E., Holstein, J. J., Hübschle, C. B. & Jayatilaka, D. (2012). Acta Cryst. A68, 435–442.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationDittrich, B., Hübschle, C. B., Luger, P. & Spackman, M. A. (2006). Acta Cryst. D62, 1325–1335.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationDolomanov, O. V., Bourhis, L. J., Gildea, R. J., Howard, J. A. K. & Puschmann, H. (2009). J. Appl. Cryst. 42, 339–341.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationDomagała, S., Fournier, B., Liebschner, D., Guillot, B. & Jelsch, C. (2012). Acta Cryst. A68, 337–351.  Web of Science CrossRef IUCr Journals Google Scholar
First citationDomagała, S., Munshi, P., Ahmed, M., Guillot, B. & Jelsch, C. (2011). Acta Cryst. B67, 63–78.  Web of Science CSD CrossRef 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 citationDoyle, A. G. & Jacobsen, E. N. (2007). Chem. Rev. 107, 5713–5743.  Web of Science CrossRef PubMed CAS Google Scholar
First citationEmsley, J. (2011). Nature's Building Blocks: an A–Z Guide to the Elements. Oxford University Press.  Google Scholar
First citationFarrugia, L. J., Kočovský, P., Senn, H. M. & Vyskočil, S. (2009). Acta Cryst. B65, 757–769.  Web of Science CSD CrossRef IUCr Journals Google Scholar
First citationFischer, R. X. & Tillmanns, E. (1988). Acta Cryst. C44, 775–776.  CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationFlaig, R., Koritsánszky, T., Janczak, J., Krane, H.-G., Morgenroth, W. & Luger, P. (1999). Angew. Chem. Int. Ed. 38, 1397–1400.  Web of Science CrossRef CAS Google Scholar
First citationFlaig, R., Koritsanszky, T., Zobel, D. & Luger, P. (1998). J. Am. Chem. Soc. 120, 2227–2238.  Web of Science CSD CrossRef CAS Google Scholar
First citationFörster, D., Messerschmidt, M. & Luger, P. L. (2005). Acta Cryst. C61, o420–o421.  Web of Science CSD CrossRef IUCr Journals Google Scholar
First citationFugel, M., Jayatilaka, D., Hupf, E., Overgaard, J., Hathwar, V. R., Macchi, P., Turner, M. J., Howard, J. A. K., Dolomanov, O. V., Puschmann, H., Iversen, B. B., Bürgi, H.-B. & Grabowsky, S. (2018). IUCrJ, 5, 32–44.  Web of Science CSD CrossRef CAS PubMed IUCr Journals Google Scholar
First citationGenoni, A., Dos Santos, L. H. R., Meyer, B. & Macchi, P. (2017). IUCrJ, 4, 136–146.  Web of Science CrossRef CAS PubMed IUCr Journals Google Scholar
First citationGildea, R. J., Bourhis, L. J., Dolomanov, O. V., Grosse-Kunstleve, R. W., Puschmann, H., Adams, P. D. & Howard, J. A. K. (2011). J. Appl. Cryst. 44, 1259–1263.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationGłowacki, E. D., Irimia-Vladu, M., Bauer, S. & Sariciftci, N. S. (2013). J. Mater. Chem. B, 1, 3742–3753.  Web of Science PubMed Google Scholar
First citationGrabowsky, S., Kalinowski, R., Weber, M., Förster, D., Paulmann, C. & Luger, P. (2009). Acta Cryst. B65, 488–501.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationGroom, C. R., Bruno, I. J., Lightfoot, M. P. & Ward, S. C. (2016). Acta Cryst. B72, 171–179.  Web of Science CrossRef IUCr Journals Google Scholar
First citationGruza, B., Chodkiewicz, M. L., Krzeszczakowska, J. & Dominiak, P. M. (2020). Acta Cryst. A76, 92–109.  Web of Science CrossRef IUCr Journals Google Scholar
First citationGuillot, R., Muzet, N., Dahaoui, S., Lecomte, C. & Jelsch, C. (2001). Acta Cryst. B57, 567–578.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationHansen, N. K. & Coppens, P. (1978). Acta Cryst. A34, 909–921.  CrossRef CAS IUCr Journals Web of Science Google Scholar
First citationHirshfeld, F. L. (1977). Theor. Chim. Acta, 44, 129–138.  CrossRef CAS Web of Science Google Scholar
First citationHolstein, J. J., Luger, P., Kalinowski, R., Mebs, S., Paulman, C. & Dittrich, B. (2010). Acta Cryst. B66, 568–577.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationHope, H. & Ottersen, T. (1978). Acta Cryst. B34, 3623–3626.  CSD CrossRef CAS IUCr Journals Web of Science Google Scholar
First citationHoser, A. A., Dominiak, P. M. & Woźniak, K. (2009). Acta Cryst. A65, 300–311.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationHübschle, C. B., Dittrich, B., Grabowsky, S., Messerschmidt, M. & Luger, P. (2008). Acta Cryst. B64, 363–374.  Web of Science CSD CrossRef IUCr Journals Google Scholar
First citationJanicki, R. & Starynowicz, P. (2010). Acta Cryst. B66, 559–567.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationJaradat, D. M. M., Mebs, S., Chęcińska, L. & Luger, P. (2007). Carbohydr. Res. 342, 1480–1489.  Web of Science CSD CrossRef PubMed CAS Google Scholar
First citationJarzembska, K. N. & Dominiak, P. M. (2012). Acta Cryst. A68, 139–147.  Web of Science CrossRef CAS 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 and A. Y. Zomaya, pp. 142–151. Berlin, Heidelberg: Springer.  Google Scholar
First citationJelsch, C., Pichon-Pesme, V., Lecomte, C. & Aubry, A. (1998). Acta Cryst. D54, 1306–1318.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationKalinowski, R., Dittrich, B., Hübschle, C. B., Paulmann, C. & Luger, P. (2007). Acta Cryst. B63, 753–767.  Web of Science CSD CrossRef IUCr Journals Google Scholar
First citationKamiński, R., Domagała, S., Jarzembska, K. N., Hoser, A. A., Sanjuan-Szklarz, W. F., Gutmann, M. J., Makal, A., Malińska, M., Bąk, J. M. & Woźniak, K. (2014). Acta Cryst. A70, 72–91.  Web of Science CSD CrossRef IUCr Journals 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 citationKumar, P., Gruza, B., Bojarowski, S. A. & Dominiak, P. M. (2019). Acta Cryst. A75, 398–408.  Web of Science CrossRef IUCr Journals Google Scholar
First citationLübben, J., Bourhis, L. J. & Dittrich, B. (2015). J. Appl. Cryst. 48, 1785–1793.  Web of Science CrossRef IUCr Journals Google Scholar
First citationLübben, J., Volkmann, C., Grabowsky, S., Edwards, A., Morgenroth, W., Fabbiani, F. P. A., Sheldrick, G. M. & Dittrich, B. (2014). Acta Cryst. A70, 309–316.  Web of Science CSD CrossRef IUCr Journals Google Scholar
First citationLübben, J., Wandtke, C. M., Hübschle, C. B., Ruf, M., Sheldrick, G. M. & Dittrich, B. (2019). Acta Cryst. A75, 50–62.  Web of Science CrossRef IUCr Journals Google Scholar
First citationMadsen, A. Ø. & Hoser, A. A. (2014). J. Appl. Cryst. 47, 2100–2104.  Web of Science CrossRef CAS IUCr Journals 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., Wieduwilt, E. K., Bergmann, J., Kleemiss, F., Meyer, B., Ruiz-López, M. F., Pal, R., Hupf, E., Beckmann, J., Piltz, R. O., Edwards, A. J., Grabowsky, S. & Genoni, A. (2019). J. Phys. Chem. Lett. 10, 6973–6982.  Web of Science CrossRef CAS PubMed Google Scholar
First citationMałecka, M., Chęcińska, L., Rybarczyk-Pirek, A., Morgenroth, W. & Paulmann, C. (2010). Acta Cryst. B66, 687–695.  Web of Science CSD CrossRef IUCr Journals Google Scholar
First citationMalinska, M. & Dauter, Z. (2016). Acta Cryst. D72, 770–779.  Web of Science CrossRef IUCr Journals Google Scholar
First citationMeents, A., Dittrich, B., Johnas, S. K. J., Thome, V. & Weckert, E. F. (2008). Acta Cryst. B64, 42–49.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationMerli, M. (2002). Z. Kristallogr. 217, 103–108.  Web of Science CrossRef CAS Google Scholar
First citationMeyer, B. & Genoni, A. (2018). J. Phys. Chem. A, 122, 8965–8981.  Web of Science CrossRef CAS PubMed Google Scholar
First citationMeyer, B., Guillot, B., Ruiz-Lopez, M. F. & Genoni, A. (2016). J. Chem. Theory Comput. 12, 1052–1067.  Web of Science CrossRef CAS PubMed Google Scholar
First citationMidgley, L., Bourhis, L. J., Dolomanov, O., Peyerimhoff, N. & Puschmann, H. (2019). arXiv: 1911.08847.  Google Scholar
First citationMüller, P., Herbst-Irmer, R., Spek, A. L., Schneider, T. R. & Sawaya, M. R. (2006). Crystal Structure Refinement: A Crystallographer's Guide to SHELXL. Oxford University Press.  Google Scholar
First citationMunshi, P. & Guru Row, T. N. (2002). Acta Cryst. B58, 1011–1017.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationMunshi, P. & Guru Row, T. N. (2006). Acta Cryst. B62, 612–626.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationMunshi, P., Thakur, T. S., Guru Row, T. N. & Desiraju, G. R. (2006). Acta Cryst. B62, 118–127.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationNassour, A., Domagala, S., Guillot, B., Leduc, T., Lecomte, C. & Jelsch, C. (2017). Acta Cryst. B73, 610–625.  Web of Science CrossRef IUCr Journals Google Scholar
First citationNassour, A., Kubicki, M., Wright, J., Borowiak, T., Dutkiewicz, G., Lecomte, C. & Jelsch, C. (2014). Acta Cryst. B70, 197–211.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationOvergaard, J., Platts, J. A. & Iversen, B. B. (2009). Acta Cryst. B65, 715–723.  Web of Science CSD CrossRef IUCr Journals Google Scholar
First citationPaul, A., Kubicki, M., Jelsch, C., Durand, P. & Lecomte, C. (2011). Acta Cryst. B67, 365–378.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationPetříček, V., Dušek, M. & Palatinus, L. (2014). Z. Kristallogr. Mater. 229, 345–352.  Google Scholar
First citationPichon-Pesme, V., Jelsch, C., Guillot, B. & Lecomte, C. (2004). Acta Cryst. A60, 204–208.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationPichon-Pesme, V., Lachekar, H., Souhassou, M. & Lecomte, C. (2000). Acta Cryst. B56, 728–737.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationPoulain, A., Wenger, E., Durand, P., Jarzembska, K. N., Kamiński, R., Fertey, P., Kubicki, M. & Lecomte, C. (2014). IUCrJ, 1, 110–118.  Web of Science CSD CrossRef CAS PubMed IUCr Journals Google Scholar
First citationRoversi, P., Merati, F., Destro, R. & Barzaghi, M. (1996). Can. J. Chem. 74, 1145–1161.  CSD CrossRef CAS Web of Science Google Scholar
First citationRuysink, A. F. J. & Vos, A. (1974). Acta Cryst. A30, 503–506.  CrossRef CAS IUCr Journals Web of Science Google Scholar
First citationScheins, S., Dittrich, B., Messerschmidt, M., Paulmann, C. & Luger, P. (2004). Acta Cryst. B60, 184–190.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationScheins, S., Messerschmidt, M. & Luger, P. (2005). Acta Cryst. B61, 443–448.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationSheldrick, G. M. (2015). Acta Cryst. C71, 3–8.  Web of Science CrossRef IUCr Journals Google Scholar
First citationŚledź, P., Kamiński, R., Chruszcz, M., Zimmerman, M. D., Minor, W. & Woźniak, K. (2010). Acta Cryst. B66, 482–492.  Web of Science CSD CrossRef IUCr Journals Google Scholar
First citationSlouf, M., Holy, A., Petříček, V. & Cisarova, I. (2002). Acta Cryst. B58, 519–529.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationSpek, A. L. (2003). J. Appl. Cryst. 36, 7–13.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationSpek, A. L. (2020). Acta Cryst. E76, 1–11.  Web of Science CrossRef IUCr Journals Google Scholar
First citationStalke, D. (2012). Electron Density and Chemical Bonding I: Experimental Charge Density Studies. Springer.  Google Scholar
First citationStewart, R. F., Bentley, J. & Goodman, B. (1975). J. Chem. Phys. 63, 3786–3793.  CrossRef CAS Web of Science 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 citationSwaminathan, S., Craven, B. M. & McMullan, R. K. (1984). Acta Cryst. B40, 300–306.  CSD CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationTaylor, M. S. & Jacobsen, E. N. (2006). Angew. Chem. Int. Ed. 45, 1520–1543.  Web of Science CrossRef CAS Google Scholar
First citationTsirelson, V. G., Stash, A. I., Potemkin, V. A., Rykounov, A. A., Shutalev, A. D., Zhurova, E. A., Zhurov, V. V., Pinkerton, A. A., Gurskaya, G. V. & Zavodnik, V. E. (2006). Acta Cryst. B62, 676–688.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationVladilo, G. & Hassanali, A. (2018). Life, 8, 1–22.  Web of Science CrossRef Google Scholar
First citationVolkov, A., Li, X., Koritsanszky, T. & Coppens, P. (2004). J. Phys. Chem. A, 108, 4283–4300.  Web of Science CrossRef CAS Google Scholar
First citationVolkov, A., Macchi, P., Farrugia, L. J., Gatti, C., Mallinson, P., Richter, T. & Koritsanszky, T. (2006). XD2006. University at Buffalo, NY, USA, University of Milan, Italy, University of Glasgow, UK, CNRISTM, Milan, Italy, and Middle Tennessee State University, TN, USA.  Google Scholar
First citationVolkov, A., Messerschmidt, M. & Coppens, P. (2007). Acta Cryst. D63, 160–170.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationWandtke, C. M., Weil, M., Simpson, J. & Dittrich, B. (2017). Acta Cryst. B73, 794–804.  Web of Science CrossRef IUCr Journals Google Scholar
First citationWhitten, A. E. & Spackman, M. A. (2006). Acta Cryst. B62, 875–888.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationWilson, A. J. C. & Geist, V. (1993). Cryst. Res. Technol. 28, 110.  CrossRef Google Scholar
First citationWoińska, M., Grabowsky, S., Dominiak, P. M., Woźniak, K. & Jayatilaka, D. (2016). Sci. Adv. 2, e1600192.  Web of Science PubMed 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
First citationZarychta, B., Pichon-Pesme, V., Guillot, B., Lecomte, C. & Jelsch, C. (2007). Acta Cryst. A63, 108–125.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationZarychta, B., Zaleski, J., Kyzioł, J., Daszkiewicz, Z. & Jelsch, C. (2011). Acta Cryst. B67, 250–262.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationZhurova, E. A. & Pinkerton, A. A. (2001). Acta Cryst. B57, 359–365.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationZhurova, E. A., Tsirelson, V. G., Zhurov, V. V., Stash, A. I. & Pinkerton, A. A. (2006). Acta Cryst. B62, 513–520.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationZhurov, V. V., Zhurova, E. A., Stash, A. I. & Pinkerton, A. A. (2011). Acta Cryst. A67, 160–173.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar

© 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