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

Journal logoSTRUCTURAL
CHEMISTRY
ISSN: 2053-2296

Two metastable high hydrates of energetic material 3,3′,5,5′-tetra­nitro-4,4′-bi­pyrazole

crossmark logo

aInorganic Chemistry Department, National Taras Shevchenko University of Kyiv, Volodymyrska Str. 64/13, 01601 Kyiv, Ukraine, and bInstitute of Inorganic Chemistry, Leipzig University, Johannisallee 29, D-04103 Leipzig, Germany
*Correspondence e-mail: dk@univ.kiev.ua

Edited by I. Oswald, University of Strathclyde, United Kingdom (Received 6 March 2024; accepted 16 April 2024; online 26 April 2024)

Poly-stoichiometry of hydrated phases is relatively uncommon for organic materials and extended libraries of such species adopting different aqua-to-substrate ratios are still rare. The kinetically controlled higher hydrates could be particularly inter­esting for their structural relationships, which presumably may imprint some features of the substrate/substrate and aqua/substrate bonding in solutions, and provide insights into the nucleation stage. Two metastable high hydrates are pre­pared by crash crystallization. The crystal structures of 3,3′,5,5′-tetra­nitro-4,4′-bi­pyrazole tetra­hydrate, C6H2N8O8·4H2O, (1), and 3,3′,5,5′-tetra­nitro-4,4′-bi­pyrazole penta­hydrate, C6H2N8O8·5H2O, (2), are intrinsically related to the previously reported anhydrate and monohydrate, while displaying natural evolution of the patterns upon progressive watering. The accumulation of the water mol­ecules causes their clustering, with the generation of one-dimensional tapes and two-dimensional layers in the genuine channel hydrates (1) and (2), respectively, versus the pocket hydrate structure of C6H2N8O8·H2O. The hydration primarily affects the pyrazole sites. It conditions the emergence of N—H⋯O and O—H⋯N hydrogen bonds, which is a destructive factor for pyrazole/pyrazole N—H⋯N hydrogen bonding. At the same time, extensive noncovalent inter­actions of the organic mol­ecules, namely, lone pair–π-hole O⋯N inter­actions of the NO2/NO2 and NO2/pyrazole types, are more com­petitive to the hydrogen bonding and the motifs of mutual organic/organic stacks remain intact with the increase in hydration. These trends agree with the results of Hirshfeld surface analysis. The contributions of the contacts involving H atoms are increased in line with the growing number of water mol­ecules, while the fraction of O⋯N/N⋯O (NO2) contacts is nearly invariant. One may postulate the significance of the lone pair–π-hole inter­actions to the aggregation of nitro species in solutions and their relevance for the sebsequent development of the solid-state patterns through nucleation.

1. Introduction

Crystallization in the form of hydrates is inherent to many types of organic materials. These com­positions are particularly significant to pharmaceutical science since the hydrated forms of active species may be more suited for applications with respect to long-term stability, solubility and drug performance (Morris, 1999[Morris, K. R. (1999). Structural Aspects of Hydrates and Solvates, in Polymorphism in Pharmaceutical Solids, ch. 4, edited by H. G Brittain. New York: Marcel Dekker.]). The general inter­est in organic hydrates is reflected in the increasing research inter­est in the crystal chemistry of such materials (Sanii et al., 2021[Sanii, R., Patyk-Kaźmierczak, E., Hua, C., Darwish, S., Pham, T., Forrest, K. A., Space, B. & Zaworotko, M. J. (2021). Cryst. Growth Des. 21, 4927-4939.]). A recently reported data-driven and topological mapping approach allows the prediction of hydrate crystal structures from an anhydrous framework (Hong et al., 2022[Hong, R. S., Mattei, A., Sheikh, A. Y. & Tuckerman, M. E. (2022). Proc. Natl Acad. Sci. USA, 119, e2204414119.]). It is believed that up to one third of organic com­pounds can afford hydrates (Stahly, 2007[Stahly, G. P. (2007). Cryst. Growth Des. 7, 1007-1026.]), while a recent analysis identified as many as 23698 unique hydrate structures within 286752 entries matching the search criteria (Werner & Swift, 2020[Werner, J. E. & Swift, J. A. (2020). CrystEngComm, 22, 7290-7297.]). The existence of hy­drated and anhydrous structures for a given com­pound is also remarkable. Only 1476 hydrates had a corresponding anhydrate in the database, although this number was possibly underestimated due to a bias of the Cambridge Structural Database (CSD; Groom et al., 2016[Groom, C. R., Bruno, I. J., Lightfoot, M. P. & Ward, S. C. (2016). Acta Cryst. B72, 171-179.]) toward structures that have only been crystallized once (Cruz-Cabeza et al., 2015[Cruz-Cabeza, A. J., Reutzel-Edens, S. M. & Bernstein, J. (2015). Chem. Soc. Rev. 44, 8619-8635.]). Such an underestimation could be even more significant when considering the existence of hydrated and anhydrous structures as a special issue of a more general case of poly-stoichiometry of hydrates. Therefore, organic com­positions with different substrate-to-aqua ratios are apparently uncommon, as revealed by the analysis of Basford (2021[Basford, P. A. (2021). Structural Relationships between Hydrates and Anhydrous Crystals: Towards Building an Understanding of Hydration in the Solid State. PhD thesis, University of Manchester, England.]). Only seven com­pounds within a set of 6575 entries adopted at least three different hydrate stoichiometries. One may suppose that the number of such systems is still limited primarily at the expense of metastable and elusive higher hydrate forms, the existence of which can be foreseen with a degree of certainty for many cases.

[Scheme 1]

Poly-stoichiometry of hydrates and the isolation of higher hydrates are particularly inter­esting in view of nucleation and crystal growth, since the crystallization of anhydrous or less hydrated forms inevitably succeeds significant desolvation of the dissolved species. Following the well-known Ostwald's rule of stages, one can intuitively assume initial crystallization of metastable highly hydrated phases. They may also be a primary outcome of crash crystallization. Such higher hydrates may inherit some features of the substrate–water inter­actions in solution, as a sufficient amount of the coordinated solvent remains an integral part of the crystal (Nangia & Desiraju, 1999[Nangia, A. & Desiraju, G. R. (1999). Chem. Commun. pp. 605-606.]). This situation is illustrated by the crystal structure of hexa­methyl­ene­tetra­mine hexa­hydrate, bearing a close resemblance to the solution environment of the substrate (Burton et al., 2009[Burton, R. C., Ferrari, E. S., Davey, R. J., Finney, J. L. & Bowron, D. T. (2009). J. Phys. Chem. B, 113, 5967-5977.]). Although crystallizations of hydrates often do not obey the rule of stages (Tian et al., 2010[Tian, F., Qu, H., Zimmermann, A., Munk, T., Jørgensen, A. C. & Rantanen, J. (2010). J. Pharm. Pharmacol. 62, 1534-1546.]), such a scenario may be more realistic for com­pounds that are most prone to forming hydrates. For example, 2,4-di­hydroxy­benzoic acid usually forms a hemihydrate, but fast cooling of hot solutions (crash crystallization) led to a very unstable monohydrate (Braun et al., 2011[Braun, D. E., Karamertzanis, P. G., Arlin, J.-B., Florence, A. J., Kahlenberg, V., Tocher, D. A., Griesser, U. J. & Price, S. L. (2011). Cryst. Growth Des. 11, 210-220.]).

The propensity of organic com­pounds for hydrate formation was associated with a (dis)balance of the available hydrogen-bond donor and acceptor sites (Desiraju, 1991[Desiraju, G. R. (1991). J. Chem. Soc. Chem. Commun. pp. 426-428.]), while a subsequent study suggested correlation rather with a total number of such functionalities within the mol­ecule (Infantes et al., 2007[Infantes, L., Fábián, L. & Motherwell, W. D. S. (2007). CrystEngComm, 9, 65-71.]). With the aim of isolating high hydrates, we have explored the behaviour of 3,3′,5,5′-tetra­nitro-4,4′-bi­pyrazole (H2Tnbpz), which is an excellent candidate for providing a set of valuable pre-requisites. First, this substrate perfectly fits the above criteria, in view of the exceptionally rich hydrogen-bonding functionality accom­panied by a striking mismatch between the numbers of available donors and acceptors, i.e. two N—H donors and ten O- and N-atom acceptors. Second, conformational flexibility of the bi­pyrazole mol­ecule is largely beneficial for polymorphism and pseudopolymorphism, which are well known for the 3,3′,5,5′-tetra­methyl-4,4′-bi­pyrazole prototype (H2Me4bpz; Boldog et al., 2003[Boldog, I., Rusanov, E. B., Sieler, J., Blaurock, S. & Domasevitch, K. V. (2003). Chem. Commun. pp. 740-741.]). An even more important functional feature is that the highly acidic N—H groups are very strong donors, while the NO2 groups, as well as the weakly basic di­nitro­pyrazole N atoms, are only poor acceptors of hydrogen bonding. In accordance with Etter's hydrogen-bonding rule (Etter, 1990[Etter, M. C. (1990). Acc. Chem. Res. 23, 120-126.]), the need for the bonding of strong proton donors and acceptors here are to be implemented preferably in the case of the substrate–aqua pair. Therefore, unlike the N—H⋯N hydrogen-bonded patterns seen for the H2Me4bpz systems (Boldog et al., 2001[Boldog, I., Rusanov, E. B., Chernega, A. N., Sieler, J. & Domasevitch, K. V. (2001). Angew. Chem. Int. Ed. 40, 3435-3438.]), rather N—H⋯OH2 inter­actions may occur in the case of H2Tnbpz. One can identify this substrate as being particularly prone either to hydrate formation or to hydrate poly-stoichiometry. Following these inputs, we succeeded in the crystallization of two metastable high hydrates and report their structures here. 3,3′,5,5′-Tetra­nitro-4,4′-bi­pyrazole and its ionic derivatives have attracted attention as perspective energetic materials, which combine sufficient performance and low sensitivity for safe applications (Gospodinov et al., 2024[Gospodinov, I., Domasevitch, K. V., Unger, C. C., Benz, M., Stierstorfer, J. & Klapötke, T. M. (2024). FirePhysChem 4, 1-9.]).

2. Experimental

2.1. Synthesis and crystallization

3,3′,5,5′-Tetra­nitro-4,4′-bi­pyrazole was pre­pared by the ni­tra­tion of 4,4′-bi­pyrazole in mixed acids (yield 92%) and was crystallized from hot water (45.0 g per 1 l) as the monohydrate (Domasevitch et al., 2019[Domasevitch, K. V., Gospodinov, I., Krautscheid, H., Klapötke, T. M. & Stierstorfer, J. (2019). New J. Chem. 43, 1305-1312.]).

To pre­pare the high hydrates, 0.0636 g (0.6 mmol) of solid Na2CO3 was added to a suspension of 0.3980 g (1.2 mmol) of H2Tnbpz·H2O in 7 ml of deionized water, and the mixture was stirred until gas evolution ceased. The mixture was then heated to 363–368 K to dissolve any solids, giving a clear light-yellow solution. Fast crystallization to a final temperature of 283 K (in ice) led to the formation of a voluminous deposit of very thin plates of H2Tnbpz·5H2O, (2), filling the entire volume of the mixture. When the mixture was allowed to stand at room temperature for several hours, this initial product lost crystallinity and gradually disappeared. Instead, a cotton-like material crystallized, representing very thin needles of H2Tnbpz·4H2O, (1). This new phase was stable when standing in the mother solution for 10–15 d, but eventually it also dissolved and com­pact blocks of the thermodynamically stable monohydrate were formed. Both (1) and (2) lost crystallinity in air within minutes. Single crystals were selected immediately after crystallization, but it was not possible to perform characterization of the materials by microanalysis.

2.2. Refinement

Crystal data, data collection and structure refinement details are summarized in Table 1[link]. All H atoms were located in difference maps and then refined with isotropic displacement parameters and with soft similarity restraints for the O—H bond lengths, which results in O—H = 0.87 (2)–0.88 (2) Å for (1) and 0.828 (17)–0.875 (17) Å for (2). For (1), the N—H bond lengths were restrained at 0.90 (3) Å. For (2), one of the water mol­ecules is unequally disordered over two closely separated positions. The partial occupancy factors of 0.86/0.14 and the separation between the contributions were refined from an isotropic model and then fixed. The O atom of the minor com­ponent was left isotropic and the H atoms were constrained as riding, with O–H = 0.85 Å.

Table 1
Experimental details

Experiments were carried out at 183 K with Cu Kα radiation using a STOE STADIVARI diffractometer. Absorption was corrected for by multi-scan methods (LANA; Koziskova et al., 2016[Koziskova, J., Hahn, F., Richter, J. & Kožíšek, J. (2016). Acta Chim. Slov. 9, 136-140.]).

  (1) (2)
Crystal data
Chemical formula C6H2N8O8·4H2O C6H2N8O8·5H2O
Mr 386.22 404.24
Crystal system, space group Orthorhombic, Pbca Monoclinic, P21/c
a, b, c (Å) 21.4196 (8), 6.1927 (2), 21.9265 (8) 11.2164 (5), 20.8114 (6), 6.6646 (3)
α, β, γ (°) 90, 90, 90 90, 90.435 (4), 90
V3) 2908.44 (18) 1555.67 (11)
Z 8 4
μ (mm−1) 1.53 1.51
Crystal size (mm) 0.10 × 0.07 × 0.05 0.09 × 0.07 × 0.03
 
Data collection
Tmin, Tmax 0.771, 0.927 0.787, 0.963
No. of measured, independent and observed [I > 2σ(I)] reflections 20365, 3101, 2195 14058, 3321, 2619
Rint 0.038 0.041
(sin θ/λ)max−1) 0.638 0.638
 
Refinement
R[F2 > 2σ(F2)], wR(F2), S 0.035, 0.092, 0.90 0.046, 0.128, 1.00
No. of reflections 3101 3321
No. of parameters 276 296
No. of restraints 32 53
H-atom treatment All H-atom parameters refined H atoms treated by a mixture of independent and constrained refinement
Δρmax, Δρmin (e Å−3) 0.20, −0.23 0.42, −0.32
Computer programs: X-AREA (Stoe & Cie, 2016[Stoe & Cie (2016). X-AREA. Stoe & Cie GmbH, Darmstadt, Germany.]), SHELXS97 (Sheldrick, 2008[Sheldrick, G. M. (2008). Acta Cryst. A64, 112-122.]), SHELXL2019 (Sheldrick, 2015[Sheldrick, G. M. (2015). Acta Cryst. C71, 3-8.]), DIAMOND (Brandenburg, 1999[Brandenburg, K. (1999). DIAMOND. Release 2.1e. Crystal Impact GbR, Bonn, Germany.]) and WinGX (Farrugia, 2012[Farrugia, L. J. (2012). J. Appl. Cryst. 45, 849-854.]).

3. Results and discussion

The anhydrate H2Tnbpz (CSD refcode PITGEH) crystallized with difficulty, using the Schlenk technique, from hot high-boiling aromatic solvents after azeotropic removal of any dissolved water (Domasevitch et al., 2019[Domasevitch, K. V., Gospodinov, I., Krautscheid, H., Klapötke, T. M. & Stierstorfer, J. (2019). New J. Chem. 43, 1305-1312.]). A strong trend for the hydration is best reflected by the formation of stable H2Tnbpz·H2O (refcode PITGIL), which dominated crystallization of the substrate from every examined common solvent under contact with ambient air. In spite of such invariant isolation of the monohydrate, further insights into this presumably rich hydrate system are possible when exploiting the special protolytic properties of H2Tnbpz. Due to the appreciable acidity of 3,5-di­nitro­pyrazole groups [pKa = 3.14 for 3,5-di­nitro­pyrazole; Janssen et al., 1973[Janssen, J. W. A. M., Kruse, C. C., Koeners, H. J. & Habraken, C. (1973). J. Heterocycl. Chem. 10, 1055-1058.]], the substrate readily forms nitro­pyrazolate salts, such as Li2(Tnbpz)(H2O)4 (Domasevitch et al., 2023[Domasevitch, K. V., Senchyk, G. A. & Krautscheid, H. (2023). Acta Cryst. E79, 657-663.]). Singly charged hydrogen bi­py­ra­zol­ate anions, [H(Tnbpz)], are also known (Gospodinov et al., 2020[Gospodinov, I., Domasevitch, K. V., Unger, C. C., Klapötke, T. M. & Stierstorfer, J. (2020). Cryst. Growth Des. 20, 755-764.]). Aqueous solutions of MI[H(Tnbpz)] may be pre­pared by neutralization with appropriate amounts of alkali metal carbonates, but the stability of such salts is rather different (Fig. 1[link]). Hydrogen bi­py­ra­zol­ates were isolated with the largest Rb and Cs cations (Domasevitch & Ponomarova, 2021[Domasevitch, K. V. & Ponomarova, V. V. (2021). Acta Cryst. E77, 1109-1115.]), but in the case of MI = Li–K, they readily dismutated to afford soluble normal salts and the excess amount of the acid was deposited in the form of H2Tnbpz·H2O. However, we have found that the outcome of the crystallization depends on the cooling rate, and fast cooling of hot solutions allows the pre­paration of the metastable highly hydrated species (1) and (2).

[Figure 1]
Figure 1
Crystallization of H2Tnbpz·nH2O as a result of prototropic dismutation of hydrogen bi­py­ra­zol­ates (M = Na). The postulated aggregation of the H(Tnpbz) anions by extensive mutual NO2/NO2 inter­actions and hydration of the pyrazole sites are well reflected by the structures of the metastable high hydrates (1) and (2).

The mol­ecular structures of the title com­pounds are shown in Figs. 2[link] and 3[link]. The main geometries of the organic frames agree well with the parameters for H2Tnbpz and its monohydrate (Domasevitch et al., 2019[Domasevitch, K. V., Gospodinov, I., Krautscheid, H., Klapötke, T. M. & Stierstorfer, J. (2019). New J. Chem. 43, 1305-1312.]). In particular, clear differentiation of the angles at the ring N atoms suggests neutral pyrazole structures with localized and immobile H atoms [N—N(H)—C = 110.35 (15)–111.34 (14)° and N(H)—N—C = 103.74 (14)–105.11 (15)°]. Certain conjugation between the nitro groups and heterocycles is indicated by the nearly flat structure of the di­nitro­pyrazole fragments, with corresponding NO2/ring dihedral angles in a range 1.36 (9)–13.52 (11)° [mean 5.09 (14) and 6.97 (12)° for (1) and (2), respectively]. The dihedral angles between two pyrazole rings, however, are significant, being 66.95 (7)° for (1) and 60.38 (7)° for (2). For H2Tnbpz·H2O, this angle was even more appreciable [78.99 (6)°]. In spite of such a twisted conformation of the organic mol­ecules, in each of the hydrates, the com­ponents afford relatively dense packing, as is indicated by packing indices of 73.6 for the monohydrate, 74.2 for (1) and 73.0 for (2), which are slightly higher than the value of 72.2 for the anhydrate structure. These values approach the upper limit of the 65–75% range expected for organic solids (Dunitz, 1995[Dunitz, J. D. (1995). X-ray Analysis and the Structure of Organic Solids, 2nd corrected reprint, pp. 106-111. Basel: Verlag Helvetica Chimica Acta.]).

[Figure 2]
Figure 2
The mol­ecular structure of H2Tnbpz tetra­hydrate, (1), showing the atom and ring labelling, and with displacement ellipsoids drawn at the 50% probability level. Dotted lines indicate hydrogen bonding.
[Figure 3]
Figure 3
The mol­ecular structure of H2Tnbpz penta­hydrate, (2), showing the atom and ring labelling, and with displacement ellipsoids drawn at the 50% probability level. One water mol­ecule is unequally disordered (0.86:0.14) and the minor contribution (O2WA) is isotropic. Dotted lines indicate hydrogen bonding.

Progressive hydration results in a very illustrative evolution of crystal patterns (Fig. 4[link]). H2Tnbpz·H2O was a genuine pocket hydrate incorporating isolated water mol­ecules, whereas the two present metastable materials are channel hydrates showing extensive aqua–aqua hydrogen bonding. An increase in the molar fraction of water mol­ecules does not lead to their uniform distribution in the lattice, but primarily causes clustering and the formation of extended hydrate networks. An immediate result of the clustering is the assembly of one-dimensional (1D) hydrate tapes in tetra­hydrate (1), but with a higher number of water mol­ecules in penta­hydrate (2), the system develops an assembly of two-dimensional (2D) hydrate layers. In both structures, the mutual bonding of water mol­ecules is directional and follows a standard geometry with typical O⋯O separations, which are 2.685 (2)–2.949 (3) Å for (1) (Table 2[link]) and 2.766 (2)–2.939 (2) Å for (2) (Table 3[link]). The longer O2W—H4⋯O5Wi hydrogen bond in (1) [O⋯O = 2.949 (3) Å and H⋯O = 2.33 (3) Å; symmetry code: (i) −x, −y + 1, −z + 1], however, exists as a shorter branch of a bifurcated inter­action with aqua and nitro acceptors.

Table 2
Hydrogen-bond geometry (Å, °) for (1)[link]

D—H⋯A D—H H⋯A DA D—H⋯A
N1—H1N⋯O1W 0.96 (2) 1.68 (2) 2.634 (2) 177 (3)
N5—H2N⋯O4Wi 0.97 (2) 1.76 (2) 2.723 (2) 172 (3)
O1W—H1⋯O3W 0.87 (2) 2.00 (2) 2.866 (2) 178 (3)
O1W—H2⋯O2W 0.87 (2) 1.97 (2) 2.771 (2) 153 (3)
O2W—H3⋯N6ii 0.87 (2) 2.10 (2) 2.942 (2) 164 (3)
O2W—H4⋯O4Wiii 0.88 (2) 2.07 (2) 2.939 (2) 168 (3)
O3W—H5⋯O3iv 0.86 (2) 2.31 (3) 3.116 (2) 156 (4)
O3W—H6⋯O2Wiii 0.88 (2) 1.90 (2) 2.766 (2) 170 (3)
O4W—H7⋯O3W 0.88 (2) 1.92 (2) 2.779 (2) 166 (3)
O4W—H8⋯O4v 0.88 (2) 2.62 (4) 3.136 (2) 118 (3)
O4W—H8⋯O6iv 0.88 (2) 2.39 (4) 3.051 (2) 132 (4)
Symmetry codes: (i) [-x, -y+1, -z+1]; (ii) [x+{\script{1\over 2}}, -y+{\script{1\over 2}}, -z+1]; (iii) [-x+{\script{1\over 2}}, y+{\script{1\over 2}}, z]; (iv) [x, -y+{\script{3\over 2}}, z+{\script{1\over 2}}]; (v) [x, -y+{\script{1\over 2}}, z+{\script{1\over 2}}].

Table 3
Hydrogen-bond geometry (Å, °) for (2)[link]

D—H⋯A D—H H⋯A DA D—H⋯A
N1—H1N⋯O1W 0.92 (3) 1.70 (3) 2.610 (2) 173 (3)
N5—H2N⋯O4Wi 0.87 (2) 1.85 (2) 2.723 (2) 179 (2)
O1W—H1⋯O2W 0.87 (2) 1.99 (2) 2.833 (2) 163 (2)
O1W—H2⋯O5Wii 0.86 (2) 1.84 (2) 2.685 (2) 170 (3)
O2W—H3⋯N6iii 0.88 (2) 2.21 (3) 3.000 (2) 151 (3)
O2W—H4⋯O5Wiv 0.83 (2) 2.33 (3) 2.949 (3) 131 (3)
O2W—H4⋯O1v 0.83 (2) 2.57 (3) 3.241 (3) 139 (3)
O3W—H5⋯N2 0.85 (2) 2.07 (2) 2.885 (2) 160 (2)
O3W—H6⋯O2W 0.83 (2) 1.99 (2) 2.802 (2) 166 (2)
O4W—H7⋯O3Wiv 0.88 (2) 1.87 (2) 2.735 (2) 173 (2)
O4W—H8⋯O3W 0.85 (2) 1.91 (2) 2.749 (2) 176 (3)
O5W—H9⋯O4W 0.85 (2) 1.94 (2) 2.784 (2) 175 (3)
O5W—H10⋯O3 0.86 (2) 2.19 (3) 2.910 (2) 141 (3)
O5W—H10⋯O8iv 0.86 (2) 2.67 (3) 3.065 (2) 109 (2)
Symmetry codes: (i) [x-1, -y+{\script{1\over 2}}, z+{\script{1\over 2}}]; (ii) [-x+1, y-{\script{1\over 2}}, -z+{\script{1\over 2}}]; (iii) [x+1, y, z-1]; (iv) [x, -y+{\script{1\over 2}}, z-{\script{1\over 2}}]; (v) [-x+1, -y, -z].
[Figure 4]
Figure 4
Evolution of the crystal patterns adopted by H2Tnbpz·nH2O as a response to the increased number of water mol­ecules. (a) The pocket hydrate PITGIL (n = 1), (b) the channel hydrate (1) (n = 4) embedding 1D aqua tapes and (c) the channel hydrate (2) (n = 5) with 2D aqua layers. For (b) and (c), the infinite aqua connectivities are orthogonal to the drawing plane.

The infinite hydrate tapes in (1), running along the b direction in the crystal, consist of distorted seven-membered rings sharing three water mol­ecules between adjacent units, which is a rare T7(3) hydrate topology according to Infantes & Motherwell (2002[Infantes, L. & Motherwell, S. (2002). Cryst­EngComm, 4, 454-461.]). The H2Tnbpz mol­ecules are integrated into the tape by double hydrogen bonding, but this scheme is relevant only for rings B, which are N—H⋯O hydrogen-bond donors and O—H⋯N hydrogen-bond acceptors (Fig. 5[link]). The former bond is particularly strong [N5⋯O4Wi = 2.723 (2) Å], whereas relative weakness of the latter inter­action [O2W⋯N6ii = 2.942 (2) Å; symmetry codes: (i) −x, −y + 1, −z + 1; (ii) x + [{1\over 2}], −y + [{1\over 2}], −z + 1] reflects the low basicity of the di­nitro­pyrazole N atom. The inter­actions of the A rings are more com­plicated and only strong N—H⋯OH2 hydrogen bonds are retained [N1⋯O1W = 2.634 (2) Å]. The resulting di­nitro­pyrazole/aqua ensemble is bound to the adjacent di­nitro­pyrazole ring as a triple donor of lone pair–π-hole inter­actions (Fig. 5[link]). In particular, the N2 atom is situated nearly above the ring centroid, with an N2⋯Cg(ring A)iii distance of 3.2492 (16) Å, while nitro–nitro O3⋯N4iii and aqua–nitro O1W⋯N3iii separations are as short as 3.031 (2) and 3.029 (2) Å, respectively [symmetry code: (iii) −x + [{1\over 2}], y + [{1\over 2}], z.]

[Figure 5]
Figure 5
(a) Projection of the structure of (1) on the ab plane, showing the topology of the aqua tape (which is highlighted with a blue strip) and the accommodation of the H2Tnbpz mol­ecules through bonding of the B rings. (b) The bonding mode of the A rings with a set of lone pair–π-hole inter­actions involving also the O1W water mol­ecule. [Symmetry codes: (i) −x, −y + 1, −z + 1; (iii) −x + [{1\over 2}], y + [{1\over 2}], z; (vi) x − [{1\over 2}], −y + [{1\over 2}], −z + 1.]

The hydrogen-bonded pattern of the H2Tnbpz mol­ecules in (2) is simpler due to a uniform function of both pyrazole rings. Similar to the B rings in the tetra­hydrate, strong N—H⋯O hydrogen bonds to water mol­ecules [N⋯O = 2.610 (2) and 2.723 (2) Å] are accom­panied by weaker O—H⋯N inter­actions with aqua donors [N⋯O = 2.885 (2) and 3.000 (2) Å]. With such different donor and acceptor bonding, each pyrazole ring is installed at the hydrate layer (Fig. 6[link]) and with two pairs of such bonds, the H2Tnbpz mol­ecules are embedded between two hydrate layers separated by 11.2164 (5) Å, which is the a parameter of the unit cell.

[Figure 6]
Figure 6
(a) Projection of the structure of (2) on the bc plane, showing the 2D aqua connectivity and how it integrates the organic mol­ecules by multiple hydrogen bonding. (b) Two H2Tnbpz mol­ecules embedded between two successive hydrate layers, with the aqua shell wrapping the hydro­philic di­nitro­pyrazole sites. [Symmetry codes: (ii) −x + 1, y − [{1\over 2}], −z + [{1\over 2}]; (iv) x, −y + [{1\over 2}], z − [{1\over 2}].]

Corrugation of the layer is conditioned by the need for the most effective inter­actions between the water mol­ecules and the hydro­philic pyrazole sites. In fact, the water mol­ecules tend to envelop di­nitro­pyrazole groups, forming a certain sphere segment around them [Fig. 6[link](b)]. In this view, the array is reminiscent of the structures of clathrate hydrates incorporating the developed water frameworks. Such a structure could be assumed as a further possible step of the pattern evolution, with a higher water-to-substrate ratio. For example, the layered structure of 3-pyrroline trihydrate is similar to (2), but in the case of the penta­hydrate, a genuine clathrate hydrate was observed (Rzepiński et al., 2019[Rzepiński, P., Nowosielska, B., Cyrański, M. K., Boese, R. & Dobrzycki, L. (2019). Cryst. Growth Des. 19, 4721-4730.]).

However, unlike many relatively small mol­ecules, such as 3-pyrroline, total encapsulation within the hydrate shell is hardly possible due to the very extensive and energetically favourable inter­actions of the H2Tnbpz mol­ecules themselves. Mutual stacking of the NO2 groups, as well as the stacking of NO2 and pyrazole groups, which are kinds of peculiar lone pair–π-hole inter­actions (Bauzá et al., 2015[Bauzá, A., Mooibroek, T. J. & Frontera, A. (2015). Chem. Commun. 51, 1491-1493.]), are particularly prevalent and such bonding may be traced either for the an­hydrate or any of the three hydrates. With a cut-off of 3.30 Å, as many as five short N⋯O contacts involving the nitro groups and also one O2N⋯OH2 contact of this type, assemble the mol­ecules into tapes along the b axis in (1) (Fig. 7[link]). We note the local motif with four doubly stacked NO2 groups and O7 atoms involved in three N⋯O contacts in the range 3.008 (2)–3.277 (2) Å (Table 4[link]). In the case of (2) (Fig. 8[link]), infinite double stacks along the c axis are generated by (N4/O3/O4)/(N4/O3/O4)vii inter­actions, with N⋯O = 3.019 (2) Å [symmetry code: (vii) x, −y + [{1\over 2}], z + [{1\over 2}]]. These separations are com­parable to the short NO2/NO2 stacking seen in Li2(Tnbpz)(H2O)4 [N⋯O = 3.0349 (15)–3.0887 (15) Å; Dom­a­sevitch et al., 2023[Domasevitch, K. V., Senchyk, G. A. & Krautscheid, H. (2023). Acta Cryst. E79, 657-663.]], although they are slightly longer than similar parameters for H2Tnbpz itself [2.911 (2) Å; Domasevitch et al., 2019[Domasevitch, K. V., Gospodinov, I., Krautscheid, H., Klapötke, T. M. & Stierstorfer, J. (2019). New J. Chem. 43, 1305-1312.]]. However, all these inter­actions are highly directional, as is indicated by the nearly orthogonal orientation of the N⋯O axes with respect to the planes of the acceptor NO2 groups (Table 4[link]). The NO2/pyrazole stacking in every case is selective to the N(H) sites, with the N⋯O separations down to 3.063 (2) Å for (1).

Table 4
Geometry of lone pair–π-hole inter­actions (Å, °) for (1) and (2)

Compound O-atom donor Group O⋯N O⋯plane φ
O⋯N (nitro)          
(1) O3 (C3/N4/O3/O4)iii 3.031 (2) 2.8316 (16) 69.10 (3)
  O6 (C6/N8/O7/O8)vii 3.002 (2) 2.8835 (18) 73.85 (3)
  O7 (C4/N7/O5/O6)ix 3.008 (2) 2.8201 (19) 69.64 (4)
  O7 (C1/N3/O1/O2)ix 3.277 (2) 3.098 (2) 70.98 (5)
  O7 (C6/N8/O7/O8)viii 3.085 (2) 2.951 (2) 73.05 (5)
  O1W (C1/N3/O1/O2)iii 3.029 (2) 2.8845 (19) 72.23 (4)
(2) O3 (C3/N4/O3/O4)iv 3.034 (2) 2.9979 (16) 81.16 (2)
  O4 (C3/N4/O3/O4)vii 3.019 (2) 2.9866 (16) 81.65 (2)
  O6 (C6/N8/O7/O8)viii 3.217 (2) 2.911 (2) 64.81 (4)
  O7 (C1/N3/O1/O2)ix 3.175 (2) 3.0653 (18) 74.89 (3)
  O1W (C1/N3/O1/O2)vi 3.132 (2) 2.9153 (19) 68.56 (4)
           
O⋯N(H) (ring)          
(1) O2 (N5/N6/C4/C5/C6)i 3.144 (2) 3.0726 (16) 77.77 (3)
  O5 (N5/N6/C4/C5/C6)xiii 3.063 (2) 2.770 (3) 64.73 (5)
(2) O1 (N5/N6/C4/C5/C6)iv 3.087 (2) 2.738 (3) 62.49 (5)
  O5 (N5/N6/C4/C5/C6)x 3.0857 (19) 2.781 (2) 64.32 (5)
  O8 (N1/N2/C1/C2/C3)ix 3.190 (2) 3.1830 (19) 86.20 (4)
Notes: O⋯plane is a distance of the O-atom donor to the mean plane of the nitro (pyrazole) group and φ is the angle of the O⋯N axis to the plane of the nitro (pyrazole) group. [Symmetry codes for (1): (iii) −x + [{1\over 2}], y + [{1\over 2}], z; (vii) x, y + 1, z; (viii) −x, −y, −z + 1; (ix) x, y − 1, z, (xiii) −x, y + [{1\over 2}], −z + [{1\over 2}]. Symmetry codes for (2): (iv) x, −y + [{1\over 2}], z?1/2 [PLEASE CLARIFY]; (vi) −x + 1, −y, −z + 1; (vii) x, −y + [{1\over 2}], z + [{1\over 2}]; (viii) x, y, z − 1; (ix) x, y, z + 1; (x) −x, −y, −z + 1.]
[Figure 7]
Figure 7
Lone pair–π-hole inter­actions of the NO2 groups in the structure of (1), with the O7 atom acting as a triple donor of such bonds. The grey circle indicates a special kind of bonding involving A rings, which is further detailed in Fig. 5[link](b). [Symmetry codes: (iii) −x + [{1\over 2}], y + [{1\over 2}], z; (vii) x, y + 1, z; (viii) −x, −y, −z + 1; (ix) x, y − 1, z.]
[Figure 8]
Figure 8
Lone pair–π-hole inter­actions of the NO2 groups in the structure of (2), arranging the molecules into 1D stacks along the c direction. [Symmetry codes: (iv) x, −y + [{1\over 2}], z − [{1\over 2}]; (vi) −x + 1, −y, −z + 1; (vii) x, −y + [{1\over 2}], z + [{1\over 2}]; (viii) x, y, z − 1; (ix) x, y, z + 1.]

Dense NO2/NO2 networks give rise also to less favourable close O⋯O contacts. For (1), they are only slightly below the sum of the van der Waals radii, e.g. O7⋯O7viii = 2.970 (3) Å [symmetry code: (viii) −x, −y, −z + 1], whereas (2) reveals an exceptionally short nonbonded contact O4⋯O6vii of 2.5963 (18) Å [symmetry code: (vii) x, −y + [{1\over 2}], z + [{1\over 2}]]. This may be com­pared with inter-polyhedral O⋯O contacts of 2.687 (2) Å in the much more robust covalent framework of NiSO4 (Wildner, 1990[Wildner, M. (1990). Z. Kristallogr. 191, 223-229.]). Such an arrangement is likely essential for the instability of the present hydrate.

One can note that the NO2/NO2 stacking tolerates well the progressively increased intensity of the substrate–aqua inter­actions in line with the increased number of the water mol­ecules (H2Tnbpz·nH2O; n = 0, 1, 4 or 5). This lone pair–π bonding is highly com­petitive to weak hydrogen bonds with nitro groups (Bauzá et al., 2017[Bauzá, A., Sharko, A. V., Senchyk, G. A., Rusanov, E. B., Frontera, A. & Domasevitch, K. V. (2017). CrystEngComm, 19, 1933-1937.]) and therefore different kinds of such stacks remain intact even in the present high hydrates. In spite of the gradually increased number of identified O—H⋯O(NO2) hydrogen bonds, which is two for the monohydrate, three for (1) and four for (2) (Table 5[link]), these inter­actions represent only the weakest hydrogen bonds in the structures [O⋯O = 2.910 (2)–3.241 (3) Å] and remain in the shadow of the apparently stronger mutual inter­actions of the nitro groups. This is contrary to the mutual hydrogen bonding of H2Tnbpz mol­ecules. Thus, four N—H⋯O/O⋯H—N hydrogen bonds per mol­ecule in H2Tnbpz itself were com­pletely eliminated even in the case of the monohydrate, whereas two N—H⋯N/N⋯H—N hydrogen bonds still retained in H2Tnbpz·H2O, but disappear irrevocably upon further hydration, in favour of stronger N—H⋯OH2 and O—H⋯N hydrogen bonds.

Table 5
Number of the identified inter­actions per single H2Tnbpz mol­ecule adopted for different hydrate environments in the structures of anhydrate PITGEH, monohydrate PITGIL and high hydrates (1) and (2)

Bond type PITGEH PITGIL (1) (2)
NH⋯N/N⋯HN 2 2
NH⋯O/O⋯HN 4
NH⋯OW 1 2 2
N⋯HOW 1 2 2
O⋯HOW 2 3 4
N⋯O/O⋯N (NO2) 8 4 7 6
N⋯O/O⋯N (ring) 10 10 4 6
O2N⋯OW 1 1 1
π⋯OW 1
O(N)⋯π/π⋯O(N) 2 2 4
N⋯NO2/O2N⋯N 2
O⋯O (NO2)a 4 4 6 3
Note: (a) the number of O⋯O contacts below 3.05 Å adopted by nitro groups.

The supra­molecular inter­actions in the H2Tnbpz systems were further assessed by Hirshfeld surface analysis (Spackman & Byrom, 1997[Spackman, M. A. & Byrom, P. G. A. (1997). Chem. Phys. Lett. 267, 215-220.]; McKinnon et al., 2004[McKinnon, J. J., Spackman, M. A. & Mitchell, A. S. (2004). Acta Cryst. B60, 627-668.]; Hirshfeld, 1977[Hirshfeld, F. L. (1977). Theor. Chim. Acta, 44, 129-138.]) performed with CrystalExplorer17 (Turner et al., 2017[Turner, M. J., McKinnon, J. J., Wolff, S. K., Grimwood, D. J., Spackman, P. R., Jayatilaka, D. & Spackman, M. A. (2017). CrystalExplorer17. University of Western Australia. http://crystalexplorer.scb.uwa.edu.au/.]). The 2D fingerprint plots for the anhydrate PITGEH [Fig. 9[link](a)] and individual H2Tnbpz mol­ecules in the monohydrate PITGIL [Fig. 9[link](b)] and the present tetra- [Fig. 9[link](c)] and penta­hydrates [Fig. 9[link](d)] indicate the prevalence of the hydrogen-bond and O⋯N/N⋯O inter­actions, while general features for the partial contributions of the different contacts are very informative for the entire series (Table 6[link]). The exceedingly large fraction of O⋯O contacts for H2Tnbpz (32.9%) indicates essential contraction upon hydration, down to 29.7% for the monohydrate, 17.9% for (1) and 21.1% for (2). However, in the latter two cases, the packing generated shorter O⋯O contacts. This is particularly relevant for (2) and is reflected by a short spike pointing to the lower left, at di + de = 2.6 Å [Fig. 6[link](d)]. The possible reasons for the ease of hydration of H2Tnbpz, as well as for the instability of the present high hydrates, may be at least in part associated with the unfavourable O⋯O contacts between the nitro groups. A similar trend for decreased contributions of O⋯N/N⋯O contacts occurs primarily at the expense of less favourable NO2/pyrazole inter­actions, since the fractions of entirely O⋯N/N⋯O (nitro) contacts themselves are almost insensitive to hydration (7.8–8.9%). The nature of the latter contacts is identical across the series and they are identified in the plots by symmetrical (about the diagonal where di = de) pairs of features. Therefore, both donor and acceptor sites of such bonds, namely, the O and N atoms of the nitro groups, appear within the individual H2Tnbpz mol­ecules. A slight shortening of these features reflects a certain weakening of the O⋯N bonds, which are 2.90 Å for H2Tnbpz, 3.00 Å for (1) and 3.05 Å for (2). A decrease in the contribution of the O⋯O and O⋯N/N⋯O (pyrazole) contacts coincides with a rapid growth of contributions from the contacts with H atoms, which are only 14.1% for H2Tnbpz and 26.7% for H2Tnbpz·H2O, but are 38.5% for (1) and even 40.8% for (2). Similar to Li2(Tnbpz)(H2O)4 (Domasevitch et al., 2023[Domasevitch, K. V., Senchyk, G. A. & Krautscheid, H. (2023). Acta Cryst. E79, 657-663.]), it may be postulated that the water hydrogen-bond donors have only a minor influence on the lone pair–π-hole inter­actions of the nitro groups, but rather they help to avoid less favourable nitro-O⋯O contacts. Unlike the NO2⋯NO2 bonds, O⋯H/H⋯O bonds for the hydrates are visualized by asymmetric fingerprint plots, clearly suggesting the prevalence of the inter­actions with water mol­ecules. The upper spikes are most sharp and their lengths indicate particularly short separations of H⋯O = 1.65 Å, corresponding to the strongest N—H⋯OW hydrogen bonds. Lower spikes indicate OH2⋯O (nitro) bonds and they are perceptibly shorter (2.10–2.20 Å) and more diffuse. That the direct N—H⋯N hydrogen bonding between the pyrazole rings is irrelevant to the high hydrates coincides with the evolution of the upper spikes in the N⋯H/H⋯N plots. A very diffuse collection of points at large di + de seen for (1) com­pletely disappears in the plot for the penta­hydrate, leaving only a single lower sharp spike, which corresponds to OH2⋯N bonds with H⋯N = 2.00 Å.

Table 6
Contributions of different kinds of the contacts (%) to the Hirshfeld surfaces of individual organic mol­ecules in anhydrate PITGEH, monohydrate PITGIL and high hydrates (1) and (2)

For the two-dimensional plots for the principal contact, see Fig. 9[link].

Contacts PITGEH PITGIL (1) (2)
All⋯O,N 80.7 69.3 56.7 54.5
All⋯H 14.1 26.7 38.5 40.8
         
O⋯H/H⋯O 20.3 27.0 37.9 35.3
N⋯H/H⋯N 5.2 8.7 5.0 8.8
C⋯H/H⋯C 0.0 0.3 0.4 1.2
N⋯O/O⋯N 27.8 21.9 21.2 19.3
C⋯O/O⋯C 11.0 8.1 8.5 10.3
C⋯N/N⋯C 0.0 0.3 2.3 0.1
O⋯O 32.9 29.7 17.9 21.1
N⋯N 2.7 3.1 3.6 0.5
H⋯H 0.0 0.8 3.2 3.4
[Figure 9]
Figure 9
2D fingerprint plots for the individual H2Tnbpz mol­ecules, reflecting the subtle structural changes seen upon progressive hydration: (a) anhydrate PITGEH; (b) H2Tnbpz·H2O PITGIL; (c) H2Tnbpz·4H2O, (1); (d) H2Tnbpz·5H2O, (2), and delineated into the principal contributions of O⋯O, O⋯N/N⋯O, O⋯H/H⋯O and N⋯H/H⋯N contacts. For details on the other contributors to the surface areas, see Table 6[link].

The inter­molecular inter­action energies were calculated using the CE B3LYP/6 31G(d,p) energy model in CrystalExplorer17 (Turner et al., 2017[Turner, M. J., McKinnon, J. J., Wolff, S. K., Grimwood, D. J., Spackman, P. R., Jayatilaka, D. & Spackman, M. A. (2017). CrystalExplorer17. University of Western Australia. http://crystalexplorer.scb.uwa.edu.au/.]), where a set of symmetry-unique inter­molecular paths were considered for the organic–organic and organic–aqua inter­actions for (1) and (2) (Table 7[link]). From the plethora of inter­molecular inter­actions, the highest energies (−49.3 to −56.1 kJ mol−1) correspond to strong N—H⋯O hydrogen bonds established by highly acidic pyrazole N—H donors and the primary contributor to the total energy is the electrostatic com­ponent (up to −83.7 kJ mol−1). These data are consistent with the energetics of azole–aqua inter­actions, which is Etot = −23.9 kJ mol−1 for relatively weak donor pyrazole itself, but −41.4 kJ mol−1 for the pair sustained by acidic penta­zole and water (Chopra et al., 2018[Chopra, N., Kaur, D. & Chopra, G. (2018). ACS Omega, 3, 12688-12702.]). In the case of hydrogen bonding to weakly basic pyrazole N-atom acceptors, Etot does not exceed −17.8 kJ mol−1. However, such inter­actions are superior to the O—H⋯O hydrogen bonds with the nitro O-atom acceptors (Table 6[link]). Beyond the particular N—H⋯O hydrogen bonds, the strongest organic–aqua inter­actions are lone pair–π-hole bonds with nitro N-atom acceptors, with Etot approaching −26.4 kJ mol−1 in response to a higher contribution of the dispersion com­ponents. These values agree with the range of −8.5 to −33.6 kJ mol−1 reported for different nitro derivatives (Daszkiewicz, 2013[Daszkiewicz, M. (2013). CrystEngComm, 15, 10427-10430.]).

Table 7
Calculated inter­action energies (kJ mol−1)

Inter­action energies were calculated employing the CE-B3LYP/6-31G(d,p) functional/basis set combination. The scale factors used to determine Etot: kele = 1.057, kpol = 0.740, kdis = 0.871 and krep = 0.618 (Mackenzie et al., 2017[Mackenzie, C. F., Spackman, P. R., Jayatilaka, D. & Spackman, M. A. (2017). IUCrJ, 4, 575-587.]).

Path Typea R (Å) Eele Epol Edis Erep Etot
(1)              
Org⋯Orgix A 6.19 −16.8 −3.2 −29.2 18.9 −33.9
Org⋯Orgi B 6.29 −16.8 −3.1 −27.0 16.8 −33.3
Org⋯Orgviii C 8.30 −2.9 −1.2 −14.1 8.6 −10.9
Org⋯Orgx D 6.86 1.3 −4.1 −28.1 19.2 −14.3
Org⋯Orgxi E 8.71 −7.4 −1.6 −8.8 4.2 −14.1
Org⋯(O1W) NH⋯OH2 5.54 −82.2 −20.1 −7.9 85.0 −56.1
Org⋯(O4W)i NH⋯OH2 5.59 −62.8 −16.1 −6.9 56.8 −49.3
Org⋯(O2W)vi OH⋯N 5.90 −28.2 −4.7 −5.9 33.4 −17.8
Org⋯(O4W)xii OH⋯O 5.69 −12.0 −1.6 −5.9 11.1 −12.1
Org⋯(O3w)xii OH⋯O 5.09 0.4 −0.8 −4.1 3.0 −1.8
Org⋯(O1W)x H2O⋯NO2 4.50 −19.1 −4.1 −10.3 10.7 −25.5
               
(2)              
Org⋯Orgiv A 6.66 −17.1 −2.7 −23.2 14.3 −31.5
Org⋯Orgx B 6.10 −18.8 −3.4 −25.6 12.5 −37.0
Org⋯Orgviii C 6.33 −14.7 −2.9 −29.6 22.3 −29.8
Org⋯Orgxi D 8.35 −2.8 −1.3 −12.0 5.1 −11.2
Org⋯(O1W) NH⋯OH2 5.56 −83.7 −20.6 −6.7 87.3 −55.7
Org⋯(O4W)i NH⋯OH2 5.61 −68.5 −17.5 −8.1 64.1 −52.9
Org⋯(O3w) OH⋯N 5.72 −22.3 −5.0 −7.2 41.8 −7.8
Org⋯(O2W)xii OH⋯N 5.88 −11.9 −4.3 −6.5 24.1 −6.6
Org⋯(O4W)xiii OH⋯O 7.10 1.1 −0.4 −2.3 1.6 −0.1
Org⋯(O5W) OH⋯O 6.77 −14.9 −2.1 −3.5 16.7 −10.0
Org⋯(O2W)v OH⋯O 6.21 −4.6 −0.8 −3.6 3.7 −6.3
Org⋯(O1W)vi H2O⋯NO2 4.43 −19.2 −3.6 −11.7 10.8 −26.4
Org⋯(O5W)vii H2O⋯π 4.29 −4.2 −1.3 −7.2 3.2 −9.6
Notes: (a) For details of the inter­action modes in (1), see Fig. 10[link], and for details of the inter­action modes in (2), see Fig. 11[link]; R is the distance between the centroids of the inter­acting mol­ecules. Symmetry codes for (1): (i) −x, −y + 1, −z + 1; (vi) x − [{1\over 2}], −y + [{1\over 2}], −z + 1; (viii) −x, −y, −z + 1; (ix) x, y − 1, z; (x) −x + [{1\over 2}], y − [{1\over 2}], z; (xi) −x, y − [{1\over 2}], −z + [{1\over 2}]; (xii) x, −y + [{3\over 2}], z − [{1\over 2}]. Symmetry codes for (2): (i) x − 1, −y + [{1\over 2}], z + [{1\over 2}]; (iv) x, −y + [{1\over 2}], z − [{1\over 2}]; (v) −x + 1, −y, −z; (vi) −x + 1, −y, −z + 1; (vii) x, −y + [{1\over 2}], z + [{1\over 2}]; (viii) x, y, z − 1; (x) −x, −y, −z + 1; (xi) −x, −y, −z + 2; (xii) x − 1, y, z + 1; (xiii) x − 1, y, z.

It is not surprising that the Edis contributors are essentially larger (up to −29.6 kJ mol−1) for the different kinds of organic–organic stacks sustained by multiple lone pair–π-hole bonds. As a result, in every case, the total inter­action energies are large and negative. This provides a realistic model for mutual stackings of H2Tnbpz, which could be rival for inter­actions with the solvent by weaker O—H⋯O hydrogen bonds or single H2O⋯NO2 bonds. This emphasis is important when involving a certain pre-nucleation effect for the crystallization of the present high hydrates (1) and (2). In the case of (1), the highest energies are associated with the formation of pairs according to Type A (Etot = −33.9 kJ mol−1), with double O⋯N (nitro) bonding, and Type B (Etot = −33.3 kJ −1), with two inversion-related O⋯N (pyrazole) inter­actions (Fig. 10[link]). Three different pairs (Types AC; Fig. 11[link]) occur for (2), with very com­parable inter­action energies up to −37.0 kJ mol−1 (Table 6[link]). The energetics of the individual O⋯N (nitro) and O⋯N (pyrazole) inter­actions are very similar, as is suggested by the values calculated for Types C and E in (1), and Type D in (2) (Etot are about −11 to −14 kJ mol−1). The latter values slightly exceed the energies of most O—H⋯O hydrogen bonds with nitro acceptors. Therefore, the extensive organic–aqua inter­actions in (1) and (2) concern predominantly the outer hydro­phylic pyrazole N/NH regions, whereas two kinds of lone pair–π-hole bonds with nitro groups provide the main motifs for the packing of shape-com­plementary H2Tnbpz mol­ecules. Also, the accumulation of additional organic–aqua inter­actions at the expense of the more favourable organic–organic bonding may be involved as a structure-destabilizing factor when com­paring tetra­hydrate (1) and very unstable penta­hydrate (2).

[Figure 10]
Figure 10
The five principal pathways (Types AE) of the inter­molecular bonding represented by different kinds of lone pair–π-hole inter­actions in the structure of (1). [Symmetry codes: (i) −x, −y + 1, −z + 1; (viii) −x, −y, −z + 1; (ix) x, y − 1, z; (x) −x + [{1\over 2}], y − [{1\over 2}], z; (xi) −x, y − [{1\over 2}], −z + [{1\over 2}].]
[Figure 11]
Figure 11
The NO2/NO2 and NO2/pyrazole N⋯O inter­actions in the structure of (2). The double inter­actions of Type D are more distal [N⋯O = 3.432 (2) Å] and weaker. [Symmetry codes: (iv) x, −y + [{1\over 2}], z − [{1\over 2}]; (viii) x, y, z − 1; (x) −x, −y, −z + 1; (xi) −x, −y, −z + 2.]

In brief, the present systems provide an attractive paradigm for the crystallization of organic hydrates. The sequence of appearance of the various forms, moving from the exclusively unstable and elusive penta­hydrate to the more stable tetra­hydrate and finally to the stable monohydrate, is entirely consistent with Ostwald's rule of stages. The library of metastable high hydrates developed in the present study is important for the understanding of nucleation and crystal growth. In terms of Steed & Steed (2015[Steed, K. M. & Steed, J. W. (2015). Chem. Rev. 115, 2895-2933.]), the metastable phases may be viewed as `fossil relics of the fastest growing crystal nuclei' and therefore they possibly reflect some preferable stacking configurations of the mol­ecules and modes of their bonding with the solvent. Since the twisted mol­ecules of 3,3′,5,5′-tetra­nitro-4,4′-bi­pyrazole, as well as the corresponding anionic species, perfectly fit one another like puzzle pieces and are prone to the generation of dense stacks, the postulated effects of pre-nucleation could be particularly significant in the present case. Therefore, this system may offer new insights not only into the area of organic hydrates, but also to the related problems of polymorphism and pseudopolymorphism.

Supporting information


Computing details top

3,3',5,5'-Tetranitro-4,4'-bipyrazole tetrahydrate (1) top
Crystal data top
C6H2N8O8·4H2ODx = 1.764 Mg m3
Mr = 386.22Cu Kα radiation, λ = 1.54186 Å
Orthorhombic, PbcaCell parameters from 14295 reflections
a = 21.4196 (8) Åθ = 4.0–79.7°
b = 6.1927 (2) ŵ = 1.53 mm1
c = 21.9265 (8) ÅT = 183 K
V = 2908.44 (18) Å3Needle, colorless
Z = 80.10 × 0.07 × 0.05 mm
F(000) = 1584
Data collection top
STOE STADIVARI
diffractometer
3101 independent reflections
Radiation source: GeniX 3D HF Cu2195 reflections with I > 2σ(I)
Graded multilayer mirror monochromatorRint = 0.038
Detector resolution: 5.81 pixels mm-1θmax = 79.8°, θmin = 4.0°
rotation method, ω scansh = 2627
Absorption correction: multi-scan
(LANA; Koziskova et al., 2016)
k = 72
Tmin = 0.771, Tmax = 0.927l = 2725
20365 measured reflections
Refinement top
Refinement on F2Secondary atom site location: difference Fourier map
Least-squares matrix: fullHydrogen site location: difference Fourier map
R[F2 > 2σ(F2)] = 0.035All H-atom parameters refined
wR(F2) = 0.092 w = 1/[σ2(Fo2) + (0.0519P)2]
where P = (Fo2 + 2Fc2)/3
S = 0.90(Δ/σ)max < 0.001
3101 reflectionsΔρmax = 0.20 e Å3
276 parametersΔρmin = 0.23 e Å3
32 restraintsExtinction correction: SHELXL2019 (Sheldrick, 2015), Fc*=kFc[1+0.001xFc2λ3/sin(2θ)]-1/4
Primary atom site location: structure-invariant direct methodsExtinction coefficient: 0.00205 (15)
Special details top

Geometry. All esds (except the esd in the dihedral angle between two l.s. planes) are estimated using the full covariance matrix. The cell esds are taken into account individually in the estimation of esds in distances, angles and torsion angles; correlations between esds in cell parameters are only used when they are defined by crystal symmetry. An approximate (isotropic) treatment of cell esds is used for estimating esds involving l.s. planes.

Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/Ueq
O10.14098 (7)0.6584 (2)0.57023 (6)0.0348 (3)
O20.06762 (7)0.4647 (2)0.52816 (6)0.0344 (3)
O30.25768 (7)0.5011 (3)0.29894 (7)0.0392 (3)
O40.16120 (7)0.4025 (3)0.28851 (6)0.0382 (3)
O50.00180 (7)0.7610 (3)0.28928 (7)0.0383 (3)
O60.09120 (7)0.8045 (2)0.32791 (6)0.0336 (3)
O70.02799 (7)0.1219 (2)0.44849 (6)0.0360 (3)
O80.11324 (7)0.0525 (2)0.47155 (7)0.0390 (3)
N10.21495 (8)0.6041 (3)0.47257 (7)0.0268 (3)
N20.24037 (7)0.5836 (3)0.41701 (7)0.0266 (3)
N30.11831 (8)0.5552 (3)0.52746 (7)0.0266 (3)
N40.20511 (8)0.4678 (3)0.31930 (7)0.0285 (3)
N50.01516 (7)0.3807 (3)0.34525 (7)0.0270 (3)
N60.01111 (7)0.1998 (2)0.37827 (7)0.0267 (3)
N70.04235 (8)0.7030 (3)0.32128 (7)0.0283 (3)
N80.06291 (8)0.0349 (3)0.44517 (7)0.0280 (3)
C10.15489 (8)0.5422 (3)0.47268 (8)0.0243 (3)
C20.13753 (8)0.4775 (3)0.41480 (7)0.0234 (3)
C30.19343 (8)0.5102 (3)0.38276 (7)0.0243 (3)
C40.03595 (8)0.5028 (3)0.35335 (7)0.0240 (3)
C50.07673 (8)0.4020 (3)0.39341 (7)0.0234 (3)
C60.04372 (8)0.2129 (3)0.40688 (7)0.0235 (3)
O1W0.28405 (7)0.7280 (2)0.56554 (6)0.0330 (3)
O2W0.37886 (7)0.5771 (3)0.64017 (7)0.0385 (3)
O3W0.21480 (8)0.7967 (3)0.67571 (7)0.0364 (3)
O4W0.12499 (8)0.4908 (3)0.70570 (7)0.0361 (3)
H1N0.2402 (13)0.644 (5)0.5067 (10)0.060 (9)*
H2N0.0520 (12)0.431 (5)0.3242 (13)0.071 (10)*
H10.2626 (13)0.751 (5)0.5985 (10)0.065 (10)*
H20.3174 (11)0.661 (5)0.5777 (13)0.055 (9)*
H30.4142 (12)0.517 (5)0.6312 (15)0.074 (11)*
H40.3835 (17)0.697 (5)0.6616 (15)0.097 (13)*
H50.2375 (17)0.851 (7)0.7040 (14)0.102 (14)*
H60.1853 (12)0.892 (4)0.6691 (13)0.067 (10)*
H70.1577 (13)0.570 (5)0.6966 (16)0.084 (12)*
H80.127 (2)0.474 (7)0.7456 (11)0.130 (17)*
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
O10.0352 (8)0.0409 (9)0.0284 (6)0.0001 (6)0.0029 (5)0.0098 (5)
O20.0297 (8)0.0396 (9)0.0340 (7)0.0074 (6)0.0051 (6)0.0027 (5)
O30.0309 (8)0.0486 (10)0.0383 (7)0.0059 (6)0.0113 (6)0.0026 (6)
O40.0352 (9)0.0482 (10)0.0312 (7)0.0031 (6)0.0007 (6)0.0073 (5)
O50.0353 (8)0.0393 (9)0.0402 (8)0.0038 (6)0.0093 (6)0.0113 (6)
O60.0324 (8)0.0314 (8)0.0370 (7)0.0056 (5)0.0016 (6)0.0063 (5)
O70.0411 (9)0.0247 (8)0.0421 (7)0.0071 (6)0.0053 (6)0.0024 (5)
O80.0345 (8)0.0344 (9)0.0481 (8)0.0002 (6)0.0106 (6)0.0082 (6)
N10.0266 (8)0.0261 (8)0.0278 (7)0.0022 (6)0.0003 (6)0.0002 (5)
N20.0240 (8)0.0267 (9)0.0292 (7)0.0024 (6)0.0008 (6)0.0019 (5)
N30.0283 (8)0.0256 (9)0.0259 (7)0.0016 (6)0.0011 (6)0.0000 (5)
N40.0272 (8)0.0292 (9)0.0289 (7)0.0003 (6)0.0034 (6)0.0003 (5)
N50.0253 (8)0.0265 (9)0.0291 (7)0.0000 (6)0.0001 (6)0.0001 (5)
N60.0251 (8)0.0228 (8)0.0321 (7)0.0006 (5)0.0007 (6)0.0003 (5)
N70.0290 (8)0.0289 (9)0.0270 (7)0.0016 (6)0.0016 (6)0.0023 (5)
N80.0296 (8)0.0246 (9)0.0299 (7)0.0004 (6)0.0026 (6)0.0001 (5)
C10.0238 (9)0.0229 (9)0.0263 (8)0.0003 (6)0.0016 (6)0.0010 (6)
C20.0251 (9)0.0203 (9)0.0248 (7)0.0007 (6)0.0011 (6)0.0017 (5)
C30.0238 (9)0.0237 (9)0.0254 (8)0.0004 (6)0.0010 (6)0.0001 (6)
C40.0257 (9)0.0224 (10)0.0238 (7)0.0005 (6)0.0007 (6)0.0008 (5)
C50.0227 (9)0.0237 (9)0.0239 (7)0.0014 (6)0.0003 (6)0.0003 (5)
C60.0234 (8)0.0212 (9)0.0261 (8)0.0000 (6)0.0003 (6)0.0003 (5)
O1W0.0296 (8)0.0383 (9)0.0312 (7)0.0018 (6)0.0023 (5)0.0008 (5)
O2W0.0308 (8)0.0362 (9)0.0484 (8)0.0016 (6)0.0038 (6)0.0076 (6)
O3W0.0344 (8)0.0396 (9)0.0353 (7)0.0019 (6)0.0040 (6)0.0034 (5)
O4W0.0351 (8)0.0388 (9)0.0344 (7)0.0054 (6)0.0034 (6)0.0024 (6)
Geometric parameters (Å, º) top
O1—N31.2343 (19)N6—C61.334 (2)
O2—N31.222 (2)N7—C41.432 (2)
O3—N41.229 (2)N8—C61.445 (2)
O4—N41.226 (2)C1—C21.382 (2)
O5—N71.231 (2)C2—C31.403 (2)
O6—N71.229 (2)C2—C51.461 (2)
O7—N81.228 (2)C4—C51.387 (2)
O8—N81.228 (2)C5—C61.399 (2)
N1—N21.340 (2)O1W—H10.87 (2)
N1—C11.342 (2)O1W—H20.87 (2)
N1—H1N0.96 (2)O2W—H30.87 (2)
N2—C31.335 (2)O2W—H40.88 (2)
N3—C11.436 (2)O3W—H50.86 (2)
N4—C31.438 (2)O3W—H60.88 (2)
N5—N61.337 (2)O4W—H70.88 (2)
N5—C41.342 (2)O4W—H80.88 (2)
N5—H2N0.97 (2)
N2—N1—C1111.34 (14)N1—C1—N3120.54 (15)
N2—N1—H1N120.4 (17)C2—C1—N3129.62 (16)
C1—N1—H1N127.9 (17)C1—C2—C3100.85 (15)
C3—N2—N1103.74 (14)C1—C2—C5128.86 (16)
O2—N3—O1125.27 (16)C3—C2—C5130.27 (15)
O2—N3—C1118.00 (15)N2—C3—C2114.24 (15)
O1—N3—C1116.74 (15)N2—C3—N4118.40 (16)
O4—N4—O3123.93 (16)C2—C3—N4127.34 (16)
O4—N4—C3117.35 (16)N5—C4—C5110.10 (16)
O3—N4—C3118.71 (16)N5—C4—N7120.07 (16)
N6—N5—C4110.35 (15)C5—C4—N7129.82 (17)
N6—N5—H2N125.7 (19)C4—C5—C6101.07 (15)
C4—N5—H2N123 (2)C4—C5—C2128.37 (16)
C6—N6—N5105.11 (15)C6—C5—C2130.56 (16)
O6—N7—O5124.88 (17)N6—C6—C5113.37 (15)
O6—N7—C4117.78 (16)N6—C6—N8118.48 (16)
O5—N7—C4117.33 (16)C5—C6—N8128.11 (16)
O7—N8—O8125.24 (17)H1—O1W—H2105 (3)
O7—N8—C6117.68 (16)H3—O2W—H4112 (3)
O8—N8—C6117.09 (16)H5—O3W—H6105 (4)
N1—C1—C2109.82 (15)H7—O4W—H8105 (3)
C1—N1—N2—C31.0 (2)N6—N5—C4—N7178.71 (14)
C4—N5—N6—C60.10 (19)O6—N7—C4—N5176.65 (16)
N2—N1—C1—C20.5 (2)O5—N7—C4—N53.7 (2)
N2—N1—C1—N3178.74 (15)O6—N7—C4—C51.7 (3)
O2—N3—C1—N1168.07 (16)O5—N7—C4—C5177.98 (18)
O1—N3—C1—N112.1 (2)N5—C4—C5—C60.02 (19)
O2—N3—C1—C214.1 (3)N7—C4—C5—C6178.48 (17)
O1—N3—C1—C2165.77 (18)N5—C4—C5—C2179.92 (16)
N1—C1—C2—C30.23 (19)N7—C4—C5—C21.5 (3)
N3—C1—C2—C3177.82 (18)C1—C2—C5—C4112.2 (2)
N1—C1—C2—C5178.66 (17)C3—C2—C5—C465.7 (3)
N3—C1—C2—C50.6 (3)C1—C2—C5—C667.8 (3)
N1—N2—C3—C21.2 (2)C3—C2—C5—C6114.2 (2)
N1—N2—C3—N4179.43 (16)N5—N6—C6—C50.1 (2)
C1—C2—C3—N20.9 (2)N5—N6—C6—N8177.89 (14)
C5—C2—C3—N2179.30 (18)C4—C5—C6—N60.04 (19)
C1—C2—C3—N4178.95 (18)C2—C5—C6—N6179.98 (17)
C5—C2—C3—N42.7 (3)C4—C5—C6—N8177.70 (17)
O4—N4—C3—N2179.33 (17)C2—C5—C6—N82.2 (3)
O3—N4—C3—N20.2 (3)O7—N8—C6—N61.3 (2)
O4—N4—C3—C22.7 (3)O8—N8—C6—N6179.11 (16)
O3—N4—C3—C2178.14 (18)O7—N8—C6—C5176.33 (17)
N6—N5—C4—C50.1 (2)O8—N8—C6—C53.3 (3)
Hydrogen-bond geometry (Å, º) top
D—H···AD—HH···AD···AD—H···A
N1—H1N···O1W0.96 (2)1.68 (2)2.634 (2)177 (3)
N5—H2N···O4Wi0.97 (2)1.76 (2)2.723 (2)172 (3)
O1W—H1···O3W0.87 (2)2.00 (2)2.866 (2)178 (3)
O1W—H2···O2W0.87 (2)1.97 (2)2.771 (2)153 (3)
O2W—H3···N6ii0.87 (2)2.10 (2)2.942 (2)164 (3)
O2W—H4···O4Wiii0.88 (2)2.07 (2)2.939 (2)168 (3)
O3W—H5···O3iv0.86 (2)2.31 (3)3.116 (2)156 (4)
O3W—H6···O2Wiii0.88 (2)1.90 (2)2.766 (2)170 (3)
O4W—H7···O3W0.88 (2)1.92 (2)2.779 (2)166 (3)
O4W—H8···O4v0.88 (2)2.62 (4)3.136 (2)118 (3)
O4W—H8···O6iv0.88 (2)2.39 (4)3.051 (2)132 (4)
Symmetry codes: (i) x, y+1, z+1; (ii) x+1/2, y+1/2, z+1; (iii) x+1/2, y+1/2, z; (iv) x, y+3/2, z+1/2; (v) x, y+1/2, z+1/2.
3,3',5,5'-Tetranitro-4,4'-bipyrazole pentahydrate (2) top
Crystal data top
C6H2N8O8·5H2OF(000) = 832
Mr = 404.24Dx = 1.726 Mg m3
Monoclinic, P21/cCu Kα radiation, λ = 1.54186 Å
a = 11.2164 (5) ÅCell parameters from 14313 reflections
b = 20.8114 (6) Åθ = 3.9–80.1°
c = 6.6646 (3) ŵ = 1.51 mm1
β = 90.435 (4)°T = 183 K
V = 1555.67 (11) Å3Needle, colorless
Z = 40.09 × 0.07 × 0.03 mm
Data collection top
STOE STADIVARI
diffractometer
3321 independent reflections
Radiation source: GeniX 3D HF Cu2619 reflections with I > 2σ(I)
Graded multilayer mirror monochromatorRint = 0.041
Detector resolution: 5.81 pixels mm-1θmax = 79.5°, θmin = 3.9°
rotation method, ω scansh = 1314
Absorption correction: multi-scan
(LANA; Koziskova et al., 2016)
k = 925
Tmin = 0.787, Tmax = 0.963l = 88
14058 measured reflections
Refinement top
Refinement on F2Primary atom site location: structure-invariant direct methods
Least-squares matrix: fullSecondary atom site location: difference Fourier map
R[F2 > 2σ(F2)] = 0.046Hydrogen site location: difference Fourier map
wR(F2) = 0.128H atoms treated by a mixture of independent and constrained refinement
S = 1.00 w = 1/[σ2(Fo2) + (0.0939P)2]
where P = (Fo2 + 2Fc2)/3
3321 reflections(Δ/σ)max < 0.001
296 parametersΔρmax = 0.42 e Å3
53 restraintsΔρmin = 0.32 e Å3
Special details top

Geometry. All esds (except the esd in the dihedral angle between two l.s. planes) are estimated using the full covariance matrix. The cell esds are taken into account individually in the estimation of esds in distances, angles and torsion angles; correlations between esds in cell parameters are only used when they are defined by crystal symmetry. An approximate (isotropic) treatment of cell esds is used for estimating esds involving l.s. planes.

Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/UeqOcc. (<1)
O10.28035 (12)0.04718 (6)0.3996 (2)0.0386 (3)
O20.13306 (12)0.01590 (6)0.5861 (2)0.0392 (3)
O30.37077 (12)0.25979 (6)0.3829 (2)0.0403 (3)
O40.19788 (12)0.25621 (6)0.5233 (2)0.0387 (3)
O50.13642 (12)0.21076 (8)0.3560 (2)0.0467 (4)
O60.03913 (12)0.18784 (7)0.2445 (2)0.0399 (3)
O70.07786 (14)0.04766 (7)1.1095 (2)0.0438 (4)
O80.23648 (12)0.07031 (7)0.9392 (2)0.0405 (3)
N10.36281 (13)0.07172 (7)0.3696 (2)0.0280 (3)
N20.38173 (13)0.13466 (7)0.3548 (2)0.0283 (3)
N30.22147 (13)0.00586 (7)0.4864 (2)0.0304 (3)
N40.28528 (13)0.23088 (7)0.4509 (2)0.0296 (3)
N50.09369 (13)0.15151 (7)0.7006 (2)0.0286 (3)
N60.05416 (14)0.11939 (7)0.8599 (2)0.0308 (3)
N70.03690 (13)0.18720 (7)0.3731 (2)0.0307 (3)
N80.12899 (14)0.07156 (7)0.9660 (2)0.0318 (3)
C10.26111 (14)0.05963 (7)0.4645 (2)0.0265 (3)
C20.20628 (15)0.11677 (7)0.5198 (2)0.0259 (3)
C30.28840 (14)0.16143 (8)0.4436 (2)0.0263 (3)
C40.00963 (14)0.15605 (8)0.5601 (3)0.0267 (3)
C50.09402 (15)0.12615 (7)0.6253 (2)0.0257 (3)
C60.05760 (15)0.10474 (8)0.8161 (2)0.0278 (3)
O1W0.54652 (12)0.00552 (6)0.2491 (2)0.0359 (3)
O3W0.61658 (12)0.18711 (6)0.2917 (2)0.0379 (3)
O2W0.71826 (17)0.08688 (8)0.0706 (3)0.0375 (4)0.86
O2WA0.7582 (16)0.0942 (9)0.123 (3)0.059 (4)*0.14
H3A0.7879300.0704530.2144960.089*0.14
H4A0.8127700.1031030.0397560.089*0.14
O4W0.67352 (12)0.31146 (6)0.1914 (2)0.0356 (3)
O5W0.49162 (15)0.38334 (7)0.3636 (3)0.0486 (4)
H30.7943 (17)0.0863 (18)0.042 (5)0.058 (10)*0.86
H40.685 (3)0.088 (2)0.041 (4)0.098 (15)*0.86
H10.603 (2)0.0235 (11)0.179 (4)0.046 (7)*
H20.540 (3)0.0326 (10)0.202 (4)0.071 (10)*
H50.5436 (16)0.1760 (12)0.283 (4)0.044 (7)*
H60.6573 (19)0.1615 (11)0.226 (4)0.043 (6)*
H70.661 (2)0.3132 (13)0.062 (3)0.053 (8)*
H80.658 (3)0.2735 (10)0.228 (4)0.063 (8)*
H90.5443 (19)0.3606 (13)0.306 (4)0.070 (9)*
H100.434 (2)0.3591 (14)0.402 (5)0.074 (10)*
H1N0.423 (2)0.0455 (13)0.328 (4)0.048 (7)*
H2N0.1678 (19)0.1639 (11)0.697 (3)0.034 (6)*
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
O10.0394 (7)0.0202 (6)0.0562 (8)0.0021 (5)0.0040 (6)0.0097 (5)
O20.0367 (7)0.0237 (7)0.0575 (8)0.0042 (5)0.0111 (6)0.0005 (5)
O30.0370 (7)0.0225 (7)0.0616 (9)0.0065 (5)0.0125 (6)0.0006 (5)
O40.0366 (7)0.0219 (6)0.0577 (8)0.0029 (5)0.0114 (6)0.0038 (5)
O50.0273 (7)0.0508 (9)0.0620 (9)0.0069 (6)0.0011 (6)0.0231 (7)
O60.0334 (7)0.0463 (8)0.0401 (7)0.0019 (6)0.0057 (5)0.0096 (6)
O70.0470 (8)0.0434 (8)0.0409 (7)0.0028 (6)0.0021 (6)0.0133 (6)
O80.0334 (7)0.0438 (8)0.0445 (7)0.0067 (6)0.0006 (5)0.0055 (6)
N10.0267 (7)0.0191 (7)0.0382 (7)0.0014 (5)0.0022 (5)0.0027 (5)
N20.0290 (7)0.0189 (7)0.0369 (7)0.0001 (5)0.0010 (6)0.0005 (5)
N30.0303 (7)0.0190 (7)0.0419 (8)0.0003 (5)0.0008 (6)0.0020 (5)
N40.0308 (7)0.0190 (7)0.0390 (7)0.0001 (5)0.0021 (6)0.0012 (5)
N50.0252 (7)0.0198 (7)0.0408 (7)0.0010 (5)0.0049 (5)0.0007 (5)
N60.0321 (8)0.0232 (7)0.0372 (7)0.0000 (5)0.0035 (6)0.0012 (5)
N70.0249 (7)0.0246 (7)0.0425 (8)0.0025 (5)0.0000 (6)0.0055 (6)
N80.0365 (8)0.0233 (7)0.0358 (7)0.0018 (6)0.0004 (6)0.0002 (5)
C10.0270 (8)0.0179 (8)0.0346 (8)0.0003 (6)0.0006 (6)0.0011 (6)
C20.0269 (8)0.0187 (7)0.0319 (7)0.0011 (6)0.0009 (6)0.0010 (5)
C30.0269 (8)0.0176 (8)0.0343 (8)0.0000 (6)0.0011 (6)0.0016 (6)
C40.0257 (8)0.0181 (8)0.0364 (8)0.0008 (5)0.0020 (6)0.0009 (6)
C50.0258 (8)0.0166 (7)0.0347 (8)0.0012 (5)0.0020 (6)0.0013 (5)
C60.0317 (8)0.0190 (7)0.0327 (8)0.0001 (6)0.0012 (6)0.0007 (6)
O1W0.0340 (7)0.0238 (7)0.0500 (7)0.0026 (5)0.0070 (6)0.0027 (5)
O3W0.0277 (7)0.0249 (7)0.0611 (9)0.0023 (5)0.0034 (6)0.0020 (5)
O2W0.0347 (9)0.0337 (9)0.0442 (10)0.0025 (7)0.0040 (8)0.0027 (7)
O4W0.0305 (7)0.0222 (7)0.0542 (8)0.0005 (5)0.0048 (5)0.0046 (5)
O5W0.0492 (9)0.0228 (7)0.0741 (11)0.0007 (6)0.0115 (8)0.0004 (6)
Geometric parameters (Å, º) top
O1—N31.2313 (19)N8—C61.451 (2)
O2—N31.216 (2)C1—C21.390 (2)
O3—N41.2221 (19)C2—C31.406 (2)
O4—N41.2161 (19)C2—C51.460 (2)
O5—N71.2236 (19)C4—C51.386 (2)
O6—N71.2134 (19)C5—C61.411 (2)
O7—N81.224 (2)O1W—H10.872 (18)
O8—N81.220 (2)O1W—H20.855 (19)
N1—N21.331 (2)O3W—H50.852 (17)
N1—C11.332 (2)O3W—H60.828 (17)
N1—H1N0.92 (3)O2W—H30.875 (19)
N2—C31.329 (2)O2W—H40.83 (2)
N3—C11.441 (2)O2WA—H3A0.8500
N4—C31.447 (2)O2WA—H4A0.8500
N5—N61.328 (2)O4W—H70.875 (17)
N5—C41.337 (2)O4W—H80.845 (19)
N5—H2N0.87 (2)O5W—H90.851 (19)
N6—C61.325 (2)O5W—H100.861 (19)
N7—C41.436 (2)
N2—N1—C1111.01 (13)C5—C4—N7130.37 (15)
N2—N1—H1N116.4 (17)C4—C5—C6100.17 (14)
C1—N1—H1N132.1 (17)C4—C5—C2129.33 (15)
C3—N2—N1104.66 (14)C6—C5—C2130.42 (15)
O2—N3—O1125.39 (15)N6—C6—C5114.03 (15)
O2—N3—C1118.13 (14)N6—C6—N8118.41 (15)
O1—N3—C1116.48 (15)C5—C6—N8127.53 (15)
O4—N4—O3124.81 (15)H1—O1W—H2105 (2)
O4—N4—C3117.78 (14)H5—O3W—H6109 (2)
O3—N4—C3117.40 (14)O2WA—O2W—H3A33.1
N6—N5—C4111.26 (14)O2WA—O2W—H4A47.1
N6—N5—H2N119.0 (15)H3A—O2W—H4A69.4
C4—N5—H2N129.6 (14)O2WA—O2W—H353 (2)
C6—N6—N5104.48 (14)H3A—O2W—H364.5
O6—N7—O5125.14 (16)H4A—O2W—H318.2
O6—N7—C4118.13 (14)O2WA—O2W—H4149 (3)
O5—N7—C4116.72 (15)H3A—O2W—H4161.5
O8—N8—O7125.10 (16)H4A—O2W—H4103.8
O8—N8—C6116.73 (15)H3—O2W—H4104 (3)
O7—N8—C6118.17 (15)O2W—O2WA—H3A124.8
N1—C1—C2110.28 (14)O2W—O2WA—H4A102.5
N1—C1—N3119.50 (14)H3A—O2WA—H4A108.4
C2—C1—N3130.14 (15)O2W—O2WA—H386 (3)
C1—C2—C3100.21 (14)H3A—O2WA—H3101.1
C1—C2—C5128.84 (15)H4A—O2WA—H328.3
C3—C2—C5130.94 (14)O2W—O2WA—H418 (2)
N2—C3—C2113.84 (14)H3A—O2WA—H4138.6
N2—C3—N4116.95 (14)H4A—O2WA—H485.7
C2—C3—N4129.21 (15)H3—O2WA—H473 (3)
N5—C4—C5110.06 (15)H7—O4W—H8107 (2)
N5—C4—N7119.54 (15)H9—O5W—H10110 (2)
C1—N1—N2—C30.03 (18)N6—N5—C4—N7177.78 (14)
C4—N5—N6—C60.81 (18)O6—N7—C4—N5176.42 (15)
N2—N1—C1—C20.29 (19)O5—N7—C4—N53.1 (2)
N2—N1—C1—N3176.78 (14)O6—N7—C4—C51.5 (3)
O2—N3—C1—N1174.75 (15)O5—N7—C4—C5178.95 (17)
O1—N3—C1—N16.0 (2)N5—C4—C5—C60.08 (18)
O2—N3—C1—C28.8 (3)N7—C4—C5—C6178.03 (17)
O1—N3—C1—C2170.46 (17)N5—C4—C5—C2177.08 (15)
N1—C1—C2—C30.46 (17)N7—C4—C5—C21.0 (3)
N3—C1—C2—C3176.22 (17)C1—C2—C5—C4117.2 (2)
N1—C1—C2—C5179.77 (16)C3—C2—C5—C461.9 (3)
N3—C1—C2—C53.1 (3)C1—C2—C5—C658.9 (3)
N1—N2—C3—C20.35 (19)C3—C2—C5—C6122.0 (2)
N1—N2—C3—N4179.72 (14)N5—N6—C6—C50.78 (19)
C1—C2—C3—N20.50 (18)N5—N6—C6—N8177.12 (14)
C5—C2—C3—N2179.79 (16)C4—C5—C6—N60.44 (19)
C1—C2—C3—N4179.78 (16)C2—C5—C6—N6176.50 (15)
C5—C2—C3—N40.9 (3)C4—C5—C6—N8177.23 (16)
O4—N4—C3—N2176.67 (15)C2—C5—C6—N85.8 (3)
O3—N4—C3—N23.1 (2)O8—N8—C6—N6165.51 (15)
O4—N4—C3—C24.1 (3)O7—N8—C6—N613.8 (2)
O3—N4—C3—C2176.20 (17)O8—N8—C6—C512.1 (3)
N6—N5—C4—C50.57 (19)O7—N8—C6—C5168.60 (16)
Hydrogen-bond geometry (Å, º) top
D—H···AD—HH···AD···AD—H···A
N1—H1N···O1W0.92 (3)1.70 (3)2.610 (2)173 (3)
N5—H2N···O4Wi0.87 (2)1.85 (2)2.723 (2)179 (2)
O1W—H1···O2W0.87 (2)1.99 (2)2.833 (2)163 (2)
O1W—H2···O5Wii0.86 (2)1.84 (2)2.685 (2)170 (3)
O2W—H3···N6iii0.88 (2)2.21 (3)3.000 (2)151 (3)
O2W—H4···O5Wiv0.83 (2)2.33 (3)2.949 (3)131 (3)
O2W—H4···O1v0.83 (2)2.57 (3)3.241 (3)139 (3)
O3W—H5···N20.85 (2)2.07 (2)2.885 (2)160 (2)
O3W—H6···O2W0.83 (2)1.99 (2)2.802 (2)166 (2)
O4W—H7···O3Wiv0.88 (2)1.87 (2)2.735 (2)173 (2)
O4W—H8···O3W0.85 (2)1.91 (2)2.749 (2)176 (3)
O5W—H9···O4W0.85 (2)1.94 (2)2.784 (2)175 (3)
O5W—H10···O30.86 (2)2.19 (3)2.910 (2)141 (3)
O5W—H10···O8iv0.86 (2)2.67 (3)3.065 (2)109 (2)
Symmetry codes: (i) x1, y+1/2, z+1/2; (ii) x+1, y1/2, z+1/2; (iii) x+1, y, z1; (iv) x, y+1/2, z1/2; (v) x+1, y, z.
Geometry of lone pair–π-hole interactions (Å, °) for (1) and (2) top
CompoundO-atom donorGroupO···NO···planeφ
O···N (nitro)
(1)O3(C3/N4/O3/O4)iii3.031 (2)2.8316 (16)69.10 (3)
O6(C6/N8/O7/O8)vii3.002 (2)2.8835 (18)73.85 (3)
O7(C4/N7/O5/O6)ix3.008 (2)2.8201 (19)69.64 (4)
O7(C1/N3/O1/O2)ix3.277 (2)3.098 (2)70.98 (5)
O7(C6/N8/O7/O8)viii3.085 (2)2.951 (2)73.05 (5)
O1w(C1/N3/O1/O2)iii3.029 (2)2.8845 (19)72.23 (4)
(2)O3(C3/N4/O3/O4)iv3.034 (2)2.9979 (16)81.16 (2)
O4(C3/N4/O3/O4)vii3.019 (2)2.9866 (16)81.65 (2)
O6(C6/N8/O7/O8)viii3.217 (2)2.911 (2)64.81 (4)
O7(C1/N3/O1/O2)ix3.175 (2)3.0653 (18)74.89 (3)
O1w(C1/N3/O1/O2)vi3.132 (2)2.9153 (19)68.56 (4)
O···N(H) (ring)
(1)O2(N5/N6/C4/C5/C6)i3.144 (2)3.0726 (16)77.77 (3)
O5(N5/N6/C4/C5/C6)xiii3.063 (2)2.770 (3)64.73 (5)
(2)O1(N5/N6/C4/C5/C6)iv3.087 (2)2.738 (3)62.49 (5)
O5(N5/N6/C4/C5/C6)x3.0857 (19)2.781 (2)64.32 (5)
O8(N1/N2/C1/C2/C3)ix3.190 (2)3.1830 (19)86.20 (4)
Notes: O···plane is a distance of the O-atom donor to the mean plane of the nitro (pyrazole) group and φ is an angle of the O···N axis to the plane of the nitro (pyrazole) group. [Symmetry codes for (1): (iii) -x+1/2, y+1/2, z; (vii) x, y+1, z; (viii) -x, -y, -z+1; (ix) x, y-1, z, (xiii) -x, y+1/2, -z+1/2. Symmetry codes for (2): (iv) x, -y+1/2, z?1/2; (vi) -x+1, -y, -z+1; (vii) x, -y+1/2, z+1/2; (viii) x, y, z-1; (ix) x, y, z+1; (x) -x, -y, -z+1.]
Number of the identified interactions per single H2Tnbpz molecule adopted for different hydrate environments in the structures of anhydrate PITGEH, monohydrate PITGIL and high hydrates (1) and (2) top
Bond typePITGEHPITGIL(1)(2)
NH···N/N···HN22--
NH···O/O···HN4---
NH···Ow-122
N···HOw-122
O···HOw-234
N···O/O···N (NO2)8476
N···O/O···N (ring)101046
O2N···Ow-111
π···Ow---1
O(N)···π/π···O(N)224-
N···NO2/O2N···N2---
O···O (NO2)a4463
Note: (a) number of O···O contacts below 3.05 Å adopted by nitro groups.
Contributions of the different kinds of the contacts (%) to the Hirshfeld surfaces of individual organic molecules in anhydrate PITGEH, monohydrate PITGIL and high hydrates (1) and (2). top
ContactsPITGEHPITGIL(1)(2)
All···O,N80.769.356.754.5
All···H14.126.738.540.8
O···H/H···O20.327.037.935.3
N···H/H···N5.28.75.08.8
C···H/H···C0.00.30.41.2
N···O/O···N27.821.921.219.3
C···O/O···C11.08.18.510.3
C···N/N···C0.00.32.30.1
O···O32.929.717.921.1
N···N2.73.13.60.5
H···H0.00.83.23.4
For the 2D plots for the principal contact, see Fig. 9.
Calculated interaction energies (kJ mol-1). top
PathTypeaR (Å)EeleEpolEdisErepEtot
(1)
Org···OrgixA6.19-16.8-3.2-29.218.9-33.9
Org···OrgiB6.29-16.8-3.1-27.016.8-33.3
Org···OrgviiiC8.30-2.9-1.2-14.18.6-10.9
Org···OrgxD6.861.3-4.1-28.119.2-14.3
Org···OrgxiE8.71-7.4-1.6-8.84.2-14.1
Org···(O1w)NH···OH25.54-82.2-20.1-7.985.0-56.1
Org···(O4w)iNH···OH25.59-62.8-16.1-6.956.8-49.3
Org···(O2w)viOH···N5.90-28.2-4.7-5.933.4-17.8
Org···(O4w)xiiOH···O5.69-12.0-1.6-5.911.1-12.1
Org···(O3w)xiiOH···O5.090.4-0.8-4.13.0-1.8
Org···(O1w)xH2O···NO24.50-19.1-4.1-10.310.7-25.5
(2)
Org···OrgivA6.66-17.1-2.7-23.214.3-31.5
Org···OrgxB6.10-18.8-3.4-25.612.5-37.0
Org···OrgviiiC6.33-14.7-2.9-29.622.3-29.8
Org···OrgxiD8.35-2.8-1.3-12.05.1-11.2
Org···(O1w)NH···OH25.56-83.7-20.6-6.787.3-55.7
Org···(O4w)iNH···OH25.61-68.5-17.5-8.164.1-52.9
Org···(O3w)OH···N5.72-22.3-5.0-7.241.8-7.8
Org···(O2w)xiiOH···N5.88-11.9-4.3-6.524.1-6.6
Org···(O4w)xiiiOH···O7.101.1-0.4-2.31.6-0.1
Org···(O5w)OH···O6.77-14.9-2.1-3.516.7-10.0
Org···(O2w)vOH···O6.21-4.6-0.8-3.63.7-6.3
Org···(O1w)viH2O···NO24.43-19.2-3.6-11.710.8-26.4
Org···(O5w)viiH2O···π4.29-4.2-1.3-7.23.2-9.6
Interaction energies were calculated employing the CE-B3LYP/6-31G(d,p) functional/basis set combination. The scale factors used to determine Etot: kele = 1.057, kpol = 0.740, kdis = 0.871, and krep = 0.618 (Mackenzie et al., 2017). (a) For details of the interaction modes in (1), see Fig. 10 and for details of the interaction modes in (2), see Fig. 11; R is a distance between centroids of the interacting molecules. Symmetry codes for (1): (i) -x, -y+1, -z+1; (vi) x-1/2, -y+1/2, -z+1; (viii) -x, -y, -z+1; (ix) x, y-1, z; (x) -x+1/2, y-1/2, z; (xi) -x, y-1/2, -z+1/2; (xii) x, -y+3/2, z-1/2; Symmetry codes for (2): (i) x-1, -y+1/2, z+1/2; (iv) x, -y+1/2, z-1/2; (v) -x+1, -y, -z; (vi) -x+1, -y, -z+1; (vii) x, -y+1/2, z+1/2; (viii) x, y, z-1; (x) -x, -y, -z+1; (xi) -x, -y, -z+2; (xii) x-1, y, z+1; (xiii) x-1, y, z.
 

Funding information

Funding for this research was provided by: Ministry of Education and Science of Ukraine (grant No. 22BF037-11).

References

First citationBasford, P. A. (2021). Structural Relationships between Hydrates and Anhydrous Crystals: Towards Building an Understanding of Hydration in the Solid State. PhD thesis, University of Manchester, England.  Google Scholar
First citationBauzá, A., Mooibroek, T. J. & Frontera, A. (2015). Chem. Commun. 51, 1491–1493.  Google Scholar
First citationBauzá, A., Sharko, A. V., Senchyk, G. A., Rusanov, E. B., Frontera, A. & Domasevitch, K. V. (2017). CrystEngComm, 19, 1933–1937.  Google Scholar
First citationBoldog, I., Rusanov, E. B., Chernega, A. N., Sieler, J. & Domasevitch, K. V. (2001). Angew. Chem. Int. Ed. 40, 3435–3438.  Web of Science CrossRef CAS Google Scholar
First citationBoldog, I., Rusanov, E. B., Sieler, J., Blaurock, S. & Domasevitch, K. V. (2003). Chem. Commun. pp. 740–741.  Web of Science CSD CrossRef Google Scholar
First citationBrandenburg, K. (1999). DIAMOND. Release 2.1e. Crystal Impact GbR, Bonn, Germany.  Google Scholar
First citationBraun, D. E., Karamertzanis, P. G., Arlin, J.-B., Florence, A. J., Kahlenberg, V., Tocher, D. A., Griesser, U. J. & Price, S. L. (2011). Cryst. Growth Des. 11, 210–220.  Web of Science CSD CrossRef CAS PubMed Google Scholar
First citationBurton, R. C., Ferrari, E. S., Davey, R. J., Finney, J. L. & Bowron, D. T. (2009). J. Phys. Chem. B, 113, 5967–5977.  CrossRef PubMed CAS Google Scholar
First citationChopra, N., Kaur, D. & Chopra, G. (2018). ACS Omega, 3, 12688–12702.  CrossRef CAS PubMed Google Scholar
First citationCruz-Cabeza, A. J., Reutzel-Edens, S. M. & Bernstein, J. (2015). Chem. Soc. Rev. 44, 8619–8635.  Web of Science CAS PubMed Google Scholar
First citationDaszkiewicz, M. (2013). CrystEngComm, 15, 10427–10430.  Web of Science CrossRef CAS Google Scholar
First citationDesiraju, G. R. (1991). J. Chem. Soc. Chem. Commun. pp. 426–428.  CrossRef Web of Science Google Scholar
First citationDomasevitch, K. V., Gospodinov, I., Krautscheid, H., Klapötke, T. M. & Stierstorfer, J. (2019). New J. Chem. 43, 1305–1312.  Web of Science CSD CrossRef CAS Google Scholar
First citationDomasevitch, K. V. & Ponomarova, V. V. (2021). Acta Cryst. E77, 1109–1115.  Web of Science CSD CrossRef IUCr Journals Google Scholar
First citationDomasevitch, K. V., Senchyk, G. A. & Krautscheid, H. (2023). Acta Cryst. E79, 657–663.  CrossRef IUCr Journals Google Scholar
First citationDunitz, J. D. (1995). X-ray Analysis and the Structure of Organic Solids, 2nd corrected reprint, pp. 106–111. Basel: Verlag Helvetica Chimica Acta.  Google Scholar
First citationEtter, M. C. (1990). Acc. Chem. Res. 23, 120–126.  CrossRef CAS Web of Science Google Scholar
First citationFarrugia, L. J. (2012). J. Appl. Cryst. 45, 849–854.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationGospodinov, I., Domasevitch, K. V., Unger, C. C., Benz, M., Stierstorfer, J. & Klapötke, T. M. (2024). FirePhysChem 4, 1–9.  CrossRef Google Scholar
First citationGospodinov, I., Domasevitch, K. V., Unger, C. C., Klapötke, T. M. & Stierstorfer, J. (2020). Cryst. Growth Des. 20, 755–764.  Web of Science CSD CrossRef CAS 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 citationHirshfeld, F. L. (1977). Theor. Chim. Acta, 44, 129–138.  CrossRef CAS Web of Science Google Scholar
First citationHong, R. S., Mattei, A., Sheikh, A. Y. & Tuckerman, M. E. (2022). Proc. Natl Acad. Sci. USA, 119, e2204414119.  CrossRef PubMed Google Scholar
First citationInfantes, L., Fábián, L. & Motherwell, W. D. S. (2007). CrystEngComm, 9, 65–71.  Web of Science CrossRef CAS Google Scholar
First citationInfantes, L. & Motherwell, S. (2002). Cryst­EngComm, 4, 454–461.  Web of Science CrossRef CAS Google Scholar
First citationJanssen, J. W. A. M., Kruse, C. C., Koeners, H. J. & Habraken, C. (1973). J. Heterocycl. Chem. 10, 1055–1058.  CrossRef CAS Web of Science Google Scholar
First citationKoziskova, J., Hahn, F., Richter, J. & Kožíšek, J. (2016). Acta Chim. Slov. 9, 136–140.  Web of Science CrossRef CAS Google Scholar
First citationMackenzie, C. F., Spackman, P. R., Jayatilaka, D. & Spackman, M. A. (2017). IUCrJ, 4, 575–587.  Web of Science CrossRef CAS PubMed IUCr Journals Google Scholar
First citationMcKinnon, J. J., Spackman, M. A. & Mitchell, A. S. (2004). Acta Cryst. B60, 627–668.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationMorris, K. R. (1999). Structural Aspects of Hydrates and Solvates, in Polymorphism in Pharmaceutical Solids, ch. 4, edited by H. G Brittain. New York: Marcel Dekker.  Google Scholar
First citationNangia, A. & Desiraju, G. R. (1999). Chem. Commun. pp. 605–606.  Web of Science CrossRef Google Scholar
First citationRzepiński, P., Nowosielska, B., Cyrański, M. K., Boese, R. & Dobrzycki, L. (2019). Cryst. Growth Des. 19, 4721–4730.  Google Scholar
First citationSanii, R., Patyk-Kaźmierczak, E., Hua, C., Darwish, S., Pham, T., Forrest, K. A., Space, B. & Zaworotko, M. J. (2021). Cryst. Growth Des. 21, 4927–4939.  CrossRef CAS PubMed Google Scholar
First citationSheldrick, G. M. (2008). Acta Cryst. A64, 112–122.  Web of Science 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 citationSpackman, M. A. & Byrom, P. G. A. (1997). Chem. Phys. Lett. 267, 215–220.  CrossRef CAS Web of Science Google Scholar
First citationStahly, G. P. (2007). Cryst. Growth Des. 7, 1007–1026.  Web of Science CrossRef CAS Google Scholar
First citationSteed, K. M. & Steed, J. W. (2015). Chem. Rev. 115, 2895–2933.  Web of Science CrossRef CAS PubMed Google Scholar
First citationStoe & Cie (2016). X-AREA. Stoe & Cie GmbH, Darmstadt, Germany.  Google Scholar
First citationTian, F., Qu, H., Zimmermann, A., Munk, T., Jørgensen, A. C. & Rantanen, J. (2010). J. Pharm. Pharmacol. 62, 1534–1546.  CrossRef CAS PubMed Google Scholar
First citationTurner, M. J., McKinnon, J. J., Wolff, S. K., Grimwood, D. J., Spackman, P. R., Jayatilaka, D. & Spackman, M. A. (2017). CrystalExplorer17. University of Western Australia. http://crystalexplorer.scb.uwa.edu.au/Google Scholar
First citationWerner, J. E. & Swift, J. A. (2020). CrystEngComm, 22, 7290–7297.  CrossRef CAS Google Scholar
First citationWildner, M. (1990). Z. Kristallogr. 191, 223–229.  CrossRef ICSD CAS Web of Science Google Scholar

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

Journal logoSTRUCTURAL
CHEMISTRY
ISSN: 2053-2296
Follow Acta Cryst. C
Sign up for e-alerts
Follow Acta Cryst. on Twitter
Follow us on facebook
Sign up for RSS feeds