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

Journal logoCRYSTALLOGRAPHIC
COMMUNICATIONS
ISSN: 2056-9890

catena-Poly[[bis­­(di­aqua­lithium)]-μ4-3,3′,5,5′-tetra­nitro-4,4′-bi­pyrazole-1,1′-diido]: a new moisture-insensitive alkali-metal energetic salt with a well-defined network structure

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 S. Parkin, University of Kentucky, USA (Received 2 June 2023; accepted 14 June 2023; online 20 June 2023)

In the structure of the title salt, [Li2(C6N8O8)(H2O)4]n, the 3,3′,5,5′-tetra­nitro-4,4′-bi­pyrazole-1,1′-diide dianion [{TNBPz}2−] is situated across the twofold axis. The distorted coordination octa­hedra around Li+ involve four short bonds with two pyrazolate N atoms and two aqua ligands [Li—N(O) = 1.999 (3)–2.090 (2) Å] and two longer contacts with nitro-O atoms [2.550 (2), 2.636 (2) Å]. When combined with μ4-{TNBPz}2−, this generates a mono-periodic polymeric structure incorporating discrete centrosymmeric [(H2O)2Li–(di­nitro­pyrazolato)2–Li(H2O)2] units. The three-dimensional stack of mutually orthogonal coordination chains is reminiscent of a Lincoln log pattern. It is influenced by conventional hydrogen bonding [O⋯O = 2.8555 (17)–3.0010 (15) Å] and multiple lone pair–π hole inter­actions of the nitro groups [N⋯O = 3.0349 (15) and 3.0887 (15) Å]. The Hirshfeld surface and two-dimensional fingerprint plots also support the significance of non-covalent bonding. Coordinative saturation and a favorable geometry at the Li+ ions, dense packing of the polymeric subconnectivities and particularly extensive inter­anion inter­actions may be involved in the stabilization of the structure. The title salt is a rare example of an energetic Li nitro­azolate, which nicely crystallizes from aqueous solution and is neither hygroscopic nor efflorescent. The TG/DTA data reveal total dehydration in the range of 330–430 K and stability of the anhydrous material up to 633–653 K.

1. Chemical context

Red-light-emitting technical or military pyrotechnics trad­itionally concern the utilization of Sr salts (Sabatini, 2018[Sabatini, J. J. (2018). Prop. Explos. Pyrotech. 43, 28-37.]). However, there is a growing inter­est for alternative red-flame colorants since strontium is potentially harmful to human health, specifically replacing calcium in bone and affecting skeletal development (Glück et al., 2017[Glück, J., Klapötke, T. M., Rusan, M., Sabatini, J. J. & Stierstorfer, J. (2017). Angew. Chem. Int. Ed. 56, 16507-16509.]). Recent works by Klapötke suggest significant potential for lithium-based systems, in particular those incorporating energetic nitro­pyrazole species (Dufter-Münster et al., 2022[Dufter-Münster, A. M. W., Harter, A. G., Klapötke, T. M., Reinhardt, E., Römer, J. & Stierstorfer, J. (2022). Eur. J. Inorg. Chem. e202101048.]; Dufter et al., 2020[Dufter, A. M. W., Klapötke, T. M., Rusan, M., Schweiger, A. & Stierstorfer, J. (2020). ChemPlusChem, 85, 2044-2050.]). The accumulation of nitro groups enhances acidity (pKa = 3.14 for 3,5-di­nitro­pyrazole vs 14.63 for the parent pyrazole; Janssen et al., 1973[Janssen, J. W. A. M., Kruse, C. C., Koeners, H. J. & Habraken, C. (1973). J. Heterocycl. Chem. 10, 1055-1058.]) for producing hydrolytically stable salts, while the incorporation of energy-rich nitro­pyrazolates contributes to oxygen balance of the formulations. In addition, the high nitro­gen content and N—N linkage within the pyrazole ring inherently facilitate the release of nitro­gen gas when burned. This meets the needs for cooling the flame for improving the color purity (Glück et al., 2017[Glück, J., Klapötke, T. M., Rusan, M., Sabatini, J. J. & Stierstorfer, J. (2017). Angew. Chem. Int. Ed. 56, 16507-16509.]).

However, most of the examined salts are still not suited for applications in spite of such valuable pre-requisites. The nitro­pyrazolates crystallize with difficulty (Drukenmüller et al., 2014[Drukenmüller, I. E., Klapötke, T. M., Morgenstern, Y., Rusan, M. & Stierstorfer, J. (2014). Z. Anorg. Allg. Chem. 640, 2139-2148.]) and their Li salts are commonly hygroscopic and deliquescent (Dufter-Münster et al., 2022[Dufter-Münster, A. M. W., Harter, A. G., Klapötke, T. M., Reinhardt, E., Römer, J. & Stierstorfer, J. (2022). Eur. J. Inorg. Chem. e202101048.]). In the present work, we address this problem with a crystal-engineering approach. The recently introduced bifunctional tecton 3,3′,5,5′-tetra­nitro-4,4′-bi­pyrazole [H2(TNBPz)] readily affords a range of salts with alkali metal ions (Domasevitch & Ponomarova, 2021[Domasevitch, K. V. & Ponomarova, V. V. (2021). Acta Cryst. E77, 1109-1115.]) and nitro­gen bases (Gospodinov et al., 2020[Gospodinov, I., Domasevitch, K. V., Unger, C. C., Klapötke, T. M. & Stierstorfer, J. (2020). Cryst. Growth Des. 20, 755-764.]) and supports either coordination or hydrogen-bonded arrays in a very predictable fashion. One can anti­cipate that the doubled nitro­pyrazolate functionality could grant the connection of the Li+ ions and generation of a relatively robust polymer, whereas the extended mol­ecular framework of {TNBPz}2− is particularly beneficial for the dense anion–anion inter­actions because of a larger contribution of dispersion forces. An appropriate set of binding sites for such inter­actions may be found with four NO2 functions, which commonly act as self-complementary donor and acceptor groups for non-covalent lone pair–π hole bonds (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.]). With this in mind, we prepared the new energetic salt catena-poly[[bis­(di­aqua­lithium)]-μ4-3,3′,5,5′-tetra­nitro-4,4′-bi­pyrazole-1,1′-diido] and report its structure here.

[Scheme 1]

2. Structural commentary

The mol­ecular structure of the title compound is shown in Fig. 1[link], with the asymmetric unit comprising one metal ion, two aqua ligands and half a mol­ecule of the organic dianion {TNBPz}2−, which is situated across the twofold axis passing through the center of the C—C bond between two pyrazole rings.

[Figure 1]
Figure 1
The mol­ecular structure of the title compound with displacement ellipsoids drawn at the 50% probability level. Dotted lines indicate distal Li—O(nitro) inter­actions. Two out of four H atoms of the aqua ligands are equally disordered over two positions (H2A, H2B and H4A, H4B). [Symmetry codes: (i) −x + 1, −y + 1, −z; (ii) −x + [{1\over 2}], −y + [{1\over 2}], z.]

The coordination around the Li+ ion may be regarded as distorted octa­hedral, with four relatively short bonds with two pyrazole-N atoms [2.086 (2) and 2.090 (2) Å], two aqua-O atoms [1.999 (3) and 2.027 (3) Å] and two elongated bonds with nitro-O atoms [2.550 (2) and 2.636 (2) Å] (Table 1[link]). A very comparable pattern for lithium 4-amino-3,5-di­nitro­pyrazolate retained only one Li—O(nitro) bond, which was slightly shorter [2.441 (4) Å; Dufter-Münster et al., 2022[Dufter-Münster, A. M. W., Harter, A. G., Klapötke, T. M., Reinhardt, E., Römer, J. & Stierstorfer, J. (2022). Eur. J. Inorg. Chem. e202101048.]]. Although the distorted octa­hedral geometries themselves are well known for Li+ (Olsher et al., 1991[Olsher, U., Izatt, R. M., Bradshaw, J. S. & Dalley, N. K. (1991). Chem. Rev. 91, 137-164.]), the spread of the bond lengths is usually narrower. For example, the citrate salt exhibits six Li—O bonds in the range of 1.998 (2)–2.222 (3) Å (Rossi et al., 1983[Rossi, M., Rickles, L. F. & Glusker, J. P. (1983). Acta Cryst. C39, 987-990.]). The exceedingly long bonds with nitro groups may be described rather as very weak ion-dipole contacts, while the remaining four shorter bonds almost perfectly match the sum of the corresponding ionic radii for 4-coordinate Li+ ions [which are Li—O = 1.94 Å and Li—N = 2.05 Å; Shannon, 1976[Shannon, R. D. (1976). Acta Cryst. A32, 751-767.]]. Nevertheless, the weak Li—O(nitro) inter­actions are presumably important for a more effective shielding of the cations against hydration when exposed to moist air. Unlike many related systems, crystals of the title compound are not hygroscopic. A second issue is the saturation of the Li+ environment with an appropriate number of aqua ligands. This is contrary to the structures of far more moisture-sensitive 3,4-, 3,5-di­nitro­pyrazolates and 4-amino-3,5-di­nitro­pyrazolate, where the di­aqua­lithium moieties were recognized as local fragments of 1:1 aqua­lithium chains –Li–[(μ-H2O)Li]n– (Dufter-Münster et al., 2022[Dufter-Münster, A. M. W., Harter, A. G., Klapötke, T. M., Reinhardt, E., Römer, J. & Stierstorfer, J. (2022). Eur. J. Inorg. Chem. e202101048.]).

Table 1
Selected geometric parameters (Å, °)

Li1—O5 1.999 (3) Li1—N2i 2.090 (2)
Li1—O6 2.027 (3) Li1—O3i 2.550 (2)
Li1—N1 2.086 (2) Li1—O1 2.636 (2)
       
O5—Li1—O6 145.63 (13) N1—Li1—O3i 170.20 (11)
O5—Li1—N1 98.49 (10) N2i—Li1—O3i 68.60 (7)
O6—Li1—N1 101.69 (11) O5—Li1—O1 80.47 (8)
O5—Li1—N2i 100.84 (11) O6—Li1—O1 81.94 (9)
O6—Li1—N2i 101.22 (11) N1—Li1—O1 67.50 (7)
N1—Li1—N2i 103.29 (10) N2i—Li1—O1 170.76 (11)
O5—Li1—O3i 78.28 (8) O3i—Li1—O1 120.49 (9)
O6—Li1—O3i 85.66 (8)    
Symmetry code: (i) [-x+1, -y+1, -z].

The geometric parameters of the {TNBPz}2− dianion are consistent with the data for neutral H2(TNBPz) (Domasevitch et al., 2019[Domasevitch, K. V., Gospodinov, I., Krautscheid, H., Klapötke, T. M. & Stierstorfer, J. (2019). New J. Chem. 43, 1305-1312.]) or singly charged species {H(TNBPz)} (Domasevitch & Ponomarova, 2021[Domasevitch, K. V. & Ponomarova, V. V. (2021). Acta Cryst. E77, 1109-1115.]). The ion adopts a twisted conformation with two pyrazole rings rotated by 54.48 (4)° (Fig. 1[link]). Although this twist is larger than 42.99 (8)° for Rb{H(TNBPz)} (Domasevitch & Ponomarova, 2021[Domasevitch, K. V. & Ponomarova, V. V. (2021). Acta Cryst. E77, 1109-1115.]), it is still unusually small, as may be compared with the even less hindered 3,3′,5,5′-tetra­methyl-4,4′-bi­pyrazole analogs (65–90°; Ponomarova et al., 2013[Ponomarova, V. V., Komarchuk, V. V., Boldog, I., Krautscheid, H. & Domasevitch, K. V. (2013). CrystEngComm, 15, 8280-8287.]). In fact, the intra­molecular inter­actions of NO2 groups, with a shortest contact N3⋯O4ii = 2.9199 (15) Å [symmetry code: (ii) −x + [{1\over 2}], −y + [{1\over 2}], z], are likely attractive, as a kind of lone pair–π hole bonding (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.]). The central C2—C2ii bond is insensitive to the protolytic effects [1.463 (2) Å vs 1.4644 (15) and 1.462 (2) Å for H2(TNBPz) and Rb{H(TNBPz)}, respectively], which indicates a lack of essential conjugation between two pyrazolate rings. Shortening of the C-NO2 bonds upon deprotonation is also minor [mean 1.4308 (16) Å vs 1.439 (2) Å for H2(TNBPz)], while a certain increase in conjugation is reflected rather by a perceptible flattening of the di­nitro­pyrazole fragments. For the latter, the NO2 groups are nearly coplanar with the ring, the two dihedral angles are 1.35 (8) and 11.66 (8)°, for N4O3O4 and N3O1O2 groups, respectively. In the case of H2(TNBPz), the twist comes to 22.8 (2)°. The most appreciable consequence of the dianionic structure is similarity of bond angles at the ring-N atoms: N2—N1—C1 = 107.11 (9)° and N1—N2—C3 = 106.93 (9)°. For the neutral di­nitro­pyrazole rings, the parameters for N- [103.9 (2)°] and NH-sites [112.0 (2)°] are clearly different (Domasevitch et al., 2019[Domasevitch, K. V., Gospodinov, I., Krautscheid, H., Klapötke, T. M. & Stierstorfer, J. (2019). New J. Chem. 43, 1305-1312.]).

3. Supra­molecular features

The title compound adopts a mono-periodic polymeric structure with the {TNBPz}2− anions acting as tetra­dentate bridging ligands. Two di­nitro­pyrazolate groups of the anions and two di­aqua­lithium fragments compose the cyclic pattern (Fig. 2[link]), which is reminiscent of the dimers in lithium 3,5-di­nitro-4-amino- and 3,4-di­nitro­pyrazolates (Dufter-Münster et al., 2022[Dufter-Münster, A. M. W., Harter, A. G., Klapötke, T. M., Reinhardt, E., Römer, J. & Stierstorfer, J. (2022). Eur. J. Inorg. Chem. e202101048.]). Unlike these monofunctional prototypes, with the C2—C2ii bond linking the two pz halves of the bi­pyrazole core, these dimers are connected into linear chains, with a distance of 9.06 Å between the centroids of the Li2(pz)2 cycles (pz is pyrazole).

[Figure 2]
Figure 2
(a) View of the coordination chains with hydrogen-bond inter­actions between pairs of symmetry-related aqua ligands that afford layers parallel to the ab plane. The Li+ ions are represented by coordination tetra­hedra, while considering only the four shortest coordination bonds. (b) Projection of the structure on the ab plane with two successive layers indicated in blue and red. (c) Packing of the coordination chains following a Lincoln log pattern. The chain nodes represent the dilithium units and the chain links are bridging {TNBPz}2− anions. [Symmetry code: (vii) −x + [{3\over 2}], −y + [{1\over 2}], z.]

Adjacent chains are linked by a set of conventional hydrogen bonds O—H⋯O, which involve either aqua or nitro-O acceptors. The geometric parameters for five types of such inter­actions are very comparable (Table 2[link]), with the range of O⋯O separations [2.8555 (17)–3.0010 (15) Å] and nearly straight angles at the H atoms [153 (2)–177 (4)°] indicating directional hydrogen bonding. Bonds of the type O6—H4A⋯O6vii [symmetry code: (vii) −x + [{3\over 2}], −y + [{1\over 2}], z; the H atom is equally disordered over two symmetry-related O atoms] are important for the connection of the chains into layers, which are parallel to the ab plane, whereas the second aqua/aqua bond O5—H⋯O6v [symmetry code: (v) −x + [{3\over 2}], y, z + [{1\over 2}]] and all three aqua/nitro hydrogen bonds actualize between the layers (Fig. 3[link]).

Table 2
Hydrogen-bond geometry (Å, °)

D—H⋯A D—H H⋯A DA D—H⋯A
O5—H1⋯O4ii 0.87 (2) 2.02 (2) 2.8818 (14) 174 (2)
O5—H2A⋯O2iii 0.87 (2) 2.13 (2) 3.0010 (15) 177 (4)
O5—H2B⋯O6iv 0.87 (2) 2.02 (2) 2.8555 (17) 163 (3)
O6—H3⋯O1v 0.82 (3) 2.24 (3) 2.9995 (17) 153 (2)
O6—H4A⋯O6vi 0.92 (4) 2.01 (4) 2.905 (2) 163 (4)
O6—H4B⋯O5vii 0.95 (5) 1.93 (5) 2.8555 (17) 164 (4)
Symmetry codes: (ii) [-x+{\script{1\over 2}}, y, z+{\script{1\over 2}}]; (iii) [-x+1, y+{\script{1\over 2}}, -z+{\script{1\over 2}}]; (iv) [-x+{\script{3\over 2}}, y, z+{\script{1\over 2}}]; (v) [x, -y+{\script{1\over 2}}, z-{\script{1\over 2}}]; (vi) [-x+{\script{3\over 2}}, -y+{\script{1\over 2}}, z]; (vii) [-x+{\script{3\over 2}}, y, z-{\script{1\over 2}}].
[Figure 3]
Figure 3
The hydrogen bonding between adjacent coordination chains, which is shown by the dotted red lines. Two successive layers are marked in blue and gray and the Li+ ions are presented as coodination tetra­hedra, while considering only the four shortest coordination bonds for clarity. The purple lines identify the directions of the coordination chains, which coincide with the crystal directions (110) and ([\overline{1}]10). [Symmetry codes: (vi) x, −y + [{1\over 2}], z − [{1\over 2}]; (vii) −x + [{3\over 2}], −y + [{1\over 2}], z; (ix) x + 1, y, z; (x) x + [{1\over 2}], y + [{1\over 2}], −z.]

The coordination chains of two successive layers are inclined, one in relation to the other, and adopt an angle of 78.9° [which is the angle between the (110) and ([\overline{1}]10) directions in the crystal]. This nearly orthogonal mutual orientation conditions a very simple packing pattern, in the form of Lincoln log-like stacks (Fig. 2[link]). The {TNBPz}2− anions are situated exactly one on the top of the other at the distances of 4.71 Å corresponding to one half of the c parameter of the unit cell. In spite of the twisted conformation of the bi­pyrazole, such stacking is geometrically favorable, with every pair of mol­ecules within the stack mutually fitting like puzzles. The resulting inter­actions are particularly extensive, with four pairs of symmetry-related short contacts N3⋯O4iii = 3.0349 (15) Å and N2⋯O2vi = 3.0887 (15) Å [symmetry codes: (iii) −x + [{1\over 2}], y, z + [{1\over 2}]; (vi) x, −y + [{1\over 2}], z − [{1\over 2}]] established by every {TNBPz}2− anion (Fig. 4[link]). For the mutually bonded nitro groups, i.e. N3O1O2 and (N4O3O4)iii, the latter is lone-pair donor and the former one is π-hole acceptor, which combine to create a very characteristic stack (Veluthaparambath et al., 2023[Veluthaparambath, R. V. P., Krishna, V., Pancharatna, P. D. & Saha, B. K. (2023). Cryst. Growth Des. 23, 442-449.]). The planes of the two groups subtend a dihedral angle of 34.16 (15)°, but the N3⋯O4iii axis is nearly orthogonal to the acceptor plane, as indicated by a slippage angle of 8.6 (2)°. The second type of inter­action of is a lone pair–π-hole bond with the di­nitro­pyrazolate ring system. Similar inter­actions are well known for electron-deficient N-heterocycles and they are most pronounced for 1,2,4,5-tetra­zines (Gural'skiy et al., 2009[Gural'skiy, I. A., Escudero, D., Frontera, A., Solntsev, P. V., Rusanov, E. B., Chernega, A. N., Krautscheid, H. & Domasevitch, K. V. (2009). Dalton Trans. pp. 2856-2864.]). In this case, the inter­planar [49.96 (10)°] and slippage angles [13.98 (15)°, with respect to the centroid of the pyrazole ring] are slightly larger. This non-covalent bonding is clearly traced in every structure adopted by H2(TNBPz) [with very short mutual nitro contacts down to N⋯O = 2.9115 (15) Å; Domasevitch et al., 2019[Domasevitch, K. V., Gospodinov, I., Krautscheid, H., Klapötke, T. M. & Stierstorfer, J. (2019). New J. Chem. 43, 1305-1312.]] and its anions (Domasevitch & Ponomarova, 2021[Domasevitch, K. V. & Ponomarova, V. V. (2021). Acta Cryst. E77, 1109-1115.]) and in fact it may be regarded as a prominent feature for the crystal chemistry of such systems. These close inter­actions of shape-complementary twisted mol­ecules contribute to the relatively high packing index of 75.8%, which is at the top 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 4]
Figure 4
Lone pair-π–hole inter­actions of the {TNBPz}2− anions. The stacking axis is indicated by the purple line and it coincides with the c-axis direction. [Symmetry codes: (iii) −x + [{1\over 2}], y, z + [{1\over 2}]; (vi) x, −y + [{1\over 2}], z − [{1\over 2}]; (xi) x, −y + [{1\over 2}], z + [{1\over 2}]; (xii) −x + [{1\over 2}], y, z − [{1\over 2}].]

4. Hirshfeld analysis

The supra­molecular inter­actions in the title structure 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.]; Spackman & McKinnon, 2002[Spackman, M. A. & McKinnon, J. J. (2002). CrystEngComm, 4, 378-392.]) 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. https://crystalexplorer.scb.uwa.edu.au/.]). The Hirshfeld surface of the individual {TNBPz}2− anion mapped over dnorm, using a fixed color scale of −0.73 (red) to 1.14 a.u. (blue), indicates a set of red spots associated with the inter­action sites (Fig. 5[link]). Two pairs of the most intense spots (−0.72 a.u.) are associated with the Li—N coordination, while the hydrogen bonding is also visualized as prominent features (−0.38 to −0.44 a.u.). The lone pair–π-hole inter­actions of NO2 groups are less visible, but they are still detectable on the surface as a set of very diffuse spots (−0.04 a.u.).

[Figure 5]
Figure 5
Hirshfeld surface of the individual {TNBPz}2− anion, mapped over dnorm (the C—H distances are normalized) in the color range −0.73 (red) to 1.14 a.u. (blue), with the red regions indicating the sites of inter­molecular inter­actions.

The two-dimensional fingerprint plots (Fig. 6[link]) are even more informative. They suggest the significance of coordination and hydrogen-bonding inter­actions, which are reflected as two sharp spikes pointing to the lower left with the shortest contacts N⋯Li = 2.1 Å and O⋯H = 2.0 Å. One can note a similar indication of N⋯Li (9.6%) and O⋯Li (4.5%) contacts, but the fraction of the latter is significantly less and the corresponding short spike is diffuse. This agrees with the weakness of the coordination bonds adopted by the nitro-O atoms. O⋯H inter­actions account for 40.1% of the entire number of contacts. This is complemented by a 10.2% contribution of N⋯H contacts, but there are no signs of any O—H⋯N bonding. The plot represents a rather diffuse collection of points with the shortest N⋯H = 2.8 Å. The large fraction of O⋯N/N⋯O and O⋯C/C⋯O contacts (in total over 20%) is a primary indicator for extensive anion–anion inter­actions. The nature of these contacts is similar and they appear in the plots as nearly symmetrical (about the diagonal where di = de) pairs of features. Therefore, either donor or acceptor sites of the bonds are found within the individual anions supporting the shortest contacts O⋯N = 3.0 Å and O⋯C = 3.1 Å. It may be postulated that the accessibility of aqua hydrogen-bond donors does not disrupt the main anion–anion inter­actions, but rather governs elimination of less favorable nitro O⋯O contacts. The total contributions of the C(N)⋯C(N,O) contacts in the title structure and in the very similar unsolvated Rb{H(TNBPz)} (Domasevitch & Ponomarova, 2021[Domasevitch, K. V. & Ponomarova, V. V. (2021). Acta Cryst. E77, 1109-1115.]) are nearly identical (33.4% and 31.9%, respectively), while the impact of hydrogen bonding is best illustrated by the pronounced contraction of the O⋯O fraction (37.4% to 11.2% in the present case). Moreover, the asymmetry of the O⋯O plot is contrary to the patterns for the O⋯N/N⋯O and O⋯C/C⋯O contacts. This witnesses the prevalence of the nitro/aqua contacts, instead of direct nitro O⋯O inter­actions.

[Figure 6]
Figure 6
Two-dimensional fingerprint plots for the individual {TNBPz}2− anion, and delineated into the principal contributions of N⋯Li, O⋯Li, O⋯H, N⋯H, O⋯N/N⋯O, O⋯C/C⋯O and O⋯O contacts. Other contributors are: C⋯N/N⋯C, 1.8%; N⋯N, 0.5% and C⋯C, 0.1% contacts.

5. Synthesis and crystallization

3,3′,5,5′-Tetra­nitro-4,4′-bi­pyrazole monohydrate [H2(TNBPz)·H2O] is readily available by nitration of 4,4′-bi­pyrazole in mixed acids (yield 92%) and subsequent crystallization from water (Domasevitch et al., 2019[Domasevitch, K. V., Gospodinov, I., Krautscheid, H., Klapötke, T. M. & Stierstorfer, J. (2019). New J. Chem. 43, 1305-1312.]).

For the preparation of the title compound, 0.294 g (7.0 mmol) of LiOH·H2O was dissolved in 10 ml of water at 333–343 K and then 1.162 g (3.5 mmol) of solid H2(TNBPz)·H2O was added with stirring. The mixture was stirred for 30 min and the resulting clear deep-yellow solution was cooled to r.t. Slow evaporation to a minimum volume over 8–10 d led to crystallization of the product as well-developed large lemon-yellow prisms. The crystals were removed and dried on a filter paper in air. The yield was 1.25 g (90%). The material shows neither signs of hygroscopy nor efflorescence when exposed to ambient air for months.

Analysis (%) calculated for C6H8Li2N8O12: C 18.10, H 2.03, N 28.15; found: C 18.45, H 1.99, N 28.42. IR (KBr, cm−1): 564 m, 636 m, 698 m, 716 m, 773 w, 853 s, 1006 m, 1021 s, 1189 w, 1284 w, 1308 s, 1321 s, 1353 s, 1381 s, 1410 s, 1484 s, 1540 s, 1641 m, 1658 m, 3460 br, 3580 br.

The FT–IR spectrum reveals a distinctive pattern, which is characteristic for hydrated nitro­pyrazolates. The peaks, which are associated with the aqua ligands, appear at 3460 and 3580 cm−1 (O—H stretching), 1641, 1658 cm−1 (bend) and 564 cm−1 (libration). The peaks for symmetric and asymmetric NO2 stretching (1351, 1381 and 1484, 1540 cm−1, respectively) are very similar to the spectra of comparable 3,5-di­nitro­pyrazole (Ravi, 2015[Ravi, P. (2015). J. Mol. Struct. 1079, 433-447.]). These double peaks originate in coupling of the NO2 vibrations with the ring motions. The intense and sharp band at 853 cm−1 is ν(C—NO2), and its shift, with respect to the band for H2(TNBPz)·H2O (839 cm−1; Domasevitch et al., 2019[Domasevitch, K. V., Gospodinov, I., Krautscheid, H., Klapötke, T. M. & Stierstorfer, J. (2019). New J. Chem. 43, 1305-1312.]), suggests a certain increase of conjugation of the nitro groups with the carrier aromatic ring upon deprotonation. For Rb{H(TNBPz)}, both these frequencies were present (839 and 852 cm−1; Domasevitch & Ponomarova, 2021[Domasevitch, K. V. & Ponomarova, V. V. (2021). Acta Cryst. E77, 1109-1115.]).

Preliminary assays for safety of the title compound and its suitability for pyrotechnic formulations were performed by thermal analysis (OZM Research DTA 552-Ex). There are two partially separated stages for nearly identical weight losses in the temperature range of 330–430 K (Fig. 7[link]), which correspond to total dehydration of the salt (in total, −18.49 mass %; −4H2O: calculated −18.09%). The anhydrous material is stable up to 633 K, with the very minor exothermic event at 597 K possibly indicating a phase transition. Exothermic decomposition proceeds above 653 K, with instantaneous loss of any remaining weight and a sharp exothermic effect at ca 700 K suggesting an explosion. For comparison, typical onset temperatures for decomposition of energetic Li nitro­pyrazolates are 400–500 K, and only 3,5-di­nitro­pyrazolate is stable up to 600 K (Dufter-Münster et al., 2022[Dufter-Münster, A. M. W., Harter, A. G., Klapötke, T. M., Reinhardt, E., Römer, J. & Stierstorfer, J. (2022). Eur. J. Inorg. Chem. e202101048.]).

[Figure 7]
Figure 7
Combined DTA (red) and TGA (blue) plots for the title compound, in the temperature range of 273–873 K (air, heating rate 5 K min−1).

6. Refinement

Crystal data, data collection and structure refinement details are summarized in Table 3[link]. The hydrogen atoms were located and then refined with isotropic thermal parameters. For both aqua ligands, one of the H atoms is equally disordered over two positions, which were refined with 0.5 partial contribution factors and with soft similarity restraints applied to O—H bond lengths [O—H = 0.82 (3)–0.95 (5) Å].

Table 3
Experimental details

Crystal data
Chemical formula [Li2(C6N8O8)(H2O)4]
Mr 398.08
Crystal system, space group Orthorhombic, Pccn
Temperature (K) 213
a, b, c (Å) 11.5094 (9), 13.9839 (8), 9.4247 (5)
V3) 1516.87 (17)
Z 4
Radiation type Mo Kα
μ (mm−1) 0.17
Crystal size (mm) 0.25 × 0.22 × 0.20
 
Data collection
Diffractometer Stoe Image plate diffraction system
No. of measured, independent and observed [I > 2σ(I)] reflections 10855, 1817, 1225
Rint 0.049
(sin θ/λ)max−1) 0.660
 
Refinement
R[F2 > 2σ(F2)], wR(F2), S 0.030, 0.072, 0.88
No. of reflections 1817
No. of parameters 151
No. of restraints 4
H-atom treatment All H-atom parameters refined
Δρmax, Δρmin (e Å−3) 0.28, −0.14
Computer programs: IPDS Software (Stoe & Cie, 2000[Stoe & Cie (2000). IPDS Software. Stoe & Cie GmbH, Darmstadt, Germany.]), SHELXS97 (Sheldrick, 2008[Sheldrick, G. M. (2008). Acta Cryst. A64, 112-122.]), SHELXL2014/7 (Sheldrick, 2015[Sheldrick, G. M. (2015). Acta Cryst. C71, 3-8.]), Diamond (Brandenburg, 1999[Brandenburg, K. (1999). DIAMOND. Crystal Impact GbR, Bonn, Germany.]) and WinGX (Farrugia, 2012[Farrugia, L. J. (2012). J. Appl. Cryst. 45, 849-854.]).

Supporting information


Computing details top

Data collection: IPDS Software (Stoe & Cie, 2000); cell refinement: IPDS Software (Stoe & Cie, 2000); data reduction: IPDS Software (Stoe & Cie, 2000); program(s) used to solve structure: SHELXS97 (Sheldrick, 2008); program(s) used to refine structure: SHELXL2014/7 (Sheldrick, 2015); molecular graphics: Diamond 2.1e (Brandenburg, 1999); software used to prepare material for publication: WinGX 1.70.01 (Farrugia, 2012).

catena-Poly[[bis(diaqualithium)]-µ4-3,3',5,5'-tetranitro-4,4'-bipyrazole-1,1'-diido] top
Crystal data top
[Li2(C6N8O8)(H2O)4]Dx = 1.743 Mg m3
Mr = 398.08Mo Kα radiation, λ = 0.71073 Å
Orthorhombic, PccnCell parameters from 8000 reflections
a = 11.5094 (9) Åθ = 2.9–28.0°
b = 13.9839 (8) ŵ = 0.17 mm1
c = 9.4247 (5) ÅT = 213 K
V = 1516.87 (17) Å3Prism, yellow
Z = 40.25 × 0.22 × 0.20 mm
F(000) = 808
Data collection top
Stoe Image plate diffraction system
diffractometer
Rint = 0.049
φ oscillation scansθmax = 28.0°, θmin = 2.9°
10855 measured reflectionsh = 1515
1817 independent reflectionsk = 1618
1225 reflections with I > 2σ(I)l = 1211
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)] = 0.030Hydrogen site location: difference Fourier map
wR(F2) = 0.072All H-atom parameters refined
S = 0.88 w = 1/[σ2(Fo2) + (0.0455P)2]
where P = (Fo2 + 2Fc2)/3
1817 reflections(Δ/σ)max < 0.001
151 parametersΔρmax = 0.28 e Å3
4 restraintsΔρmin = 0.14 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)
Li10.62380 (18)0.42289 (16)0.0859 (2)0.0258 (5)
O10.57465 (8)0.25173 (7)0.18668 (12)0.0401 (3)
O20.42324 (9)0.16320 (7)0.21808 (12)0.0357 (3)
O30.17428 (9)0.50551 (7)0.11938 (11)0.0341 (3)
O40.08500 (8)0.37055 (7)0.08988 (11)0.0327 (3)
O50.62403 (9)0.45230 (8)0.29341 (11)0.0336 (3)
O60.69704 (11)0.34427 (9)0.07111 (14)0.0361 (3)
N10.44846 (8)0.38814 (7)0.06362 (11)0.0196 (2)
N20.36602 (8)0.43947 (7)0.00029 (12)0.0191 (2)
N30.47107 (9)0.23391 (7)0.16949 (12)0.0236 (3)
N40.17120 (9)0.42244 (8)0.07765 (11)0.0212 (2)
C10.40259 (10)0.30164 (9)0.09198 (13)0.0187 (3)
C20.28824 (10)0.29180 (9)0.04627 (13)0.0176 (3)
C30.27171 (10)0.38324 (8)0.01021 (13)0.0178 (3)
H10.5584 (14)0.4317 (13)0.327 (2)0.071 (7)*
H2A0.608 (3)0.5133 (13)0.289 (4)0.064 (13)*0.5
H2B0.671 (2)0.409 (2)0.326 (4)0.047 (10)*0.5
H30.651 (2)0.3349 (17)0.136 (3)0.089 (9)*
H4A0.724 (4)0.284 (3)0.053 (4)0.062 (13)*0.5
H4B0.749 (4)0.390 (3)0.108 (5)0.081 (15)*0.5
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
Li10.0233 (10)0.0256 (12)0.0285 (13)0.0004 (9)0.0005 (9)0.0042 (10)
O10.0233 (5)0.0431 (6)0.0540 (7)0.0006 (4)0.0134 (5)0.0113 (5)
O20.0439 (6)0.0233 (5)0.0397 (6)0.0057 (4)0.0055 (5)0.0132 (4)
O30.0329 (5)0.0264 (5)0.0431 (7)0.0014 (4)0.0059 (4)0.0133 (5)
O40.0219 (5)0.0358 (6)0.0406 (6)0.0075 (4)0.0109 (4)0.0059 (5)
O50.0371 (6)0.0374 (7)0.0264 (6)0.0090 (5)0.0017 (5)0.0036 (5)
O60.0336 (6)0.0380 (7)0.0369 (7)0.0057 (5)0.0007 (5)0.0095 (5)
N10.0195 (5)0.0176 (5)0.0218 (6)0.0027 (4)0.0024 (4)0.0020 (4)
N20.0190 (5)0.0186 (5)0.0197 (5)0.0032 (4)0.0003 (4)0.0014 (4)
N30.0257 (6)0.0221 (6)0.0229 (6)0.0016 (4)0.0038 (5)0.0016 (5)
N40.0207 (5)0.0226 (6)0.0204 (6)0.0010 (4)0.0013 (4)0.0016 (5)
C10.0187 (6)0.0178 (6)0.0195 (6)0.0028 (5)0.0011 (5)0.0014 (5)
C20.0184 (6)0.0178 (6)0.0167 (6)0.0026 (5)0.0008 (5)0.0013 (5)
C30.0177 (6)0.0183 (6)0.0174 (6)0.0021 (5)0.0006 (5)0.0007 (5)
Geometric parameters (Å, º) top
Li1—O51.999 (3)O5—H2B0.866 (17)
Li1—O62.027 (3)O6—H30.82 (3)
Li1—N12.086 (2)O6—H4A0.92 (4)
Li1—N2i2.090 (2)O6—H4B0.95 (5)
Li1—O3i2.550 (2)N1—N21.3336 (14)
Li1—O12.636 (2)N1—C11.3465 (16)
O1—N31.2287 (14)N2—C31.3435 (15)
O2—N31.2209 (14)N2—Li1i2.090 (2)
O3—N41.2270 (14)N3—C11.4324 (16)
O3—Li1i2.550 (2)N4—C31.4292 (16)
O4—N41.2346 (13)C1—C21.3917 (16)
O5—H10.869 (15)C2—C31.3981 (17)
O5—H2A0.874 (17)C2—C2ii1.463 (2)
O5—Li1—O6145.63 (13)H3—O6—H4A102 (3)
O5—Li1—N198.49 (10)Li1—O6—H4B99 (3)
O6—Li1—N1101.69 (11)H3—O6—H4B105 (3)
O5—Li1—N2i100.84 (11)N2—N1—C1107.11 (9)
O6—Li1—N2i101.22 (11)N2—N1—Li1127.46 (10)
N1—Li1—N2i103.29 (10)C1—N1—Li1124.64 (10)
O5—Li1—O3i78.28 (8)N1—N2—C3106.93 (9)
O6—Li1—O3i85.66 (8)N1—N2—Li1i129.06 (9)
N1—Li1—O3i170.20 (11)C3—N2—Li1i123.87 (10)
N2i—Li1—O3i68.60 (7)O2—N3—O1123.53 (11)
O5—Li1—O180.47 (8)O2—N3—C1118.60 (10)
O6—Li1—O181.94 (9)O1—N3—C1117.84 (10)
N1—Li1—O167.50 (7)O3—N4—O4123.35 (11)
N2i—Li1—O1170.76 (11)O3—N4—C3118.83 (10)
O3i—Li1—O1120.49 (9)O4—N4—C3117.81 (10)
N3—O1—Li1110.17 (9)N1—C1—C2113.46 (11)
N4—O3—Li1i111.00 (9)N1—C1—N3118.65 (10)
Li1—O5—H1106.9 (14)C2—C1—N3127.80 (11)
Li1—O5—H2A99 (3)C1—C2—C398.98 (10)
H1—O5—H2A99.0 (19)C1—C2—C2ii130.38 (13)
Li1—O5—H2B102 (2)C3—C2—C2ii130.46 (12)
H1—O5—H2B100.4 (19)N2—C3—C2113.51 (10)
Li1—O6—H3111.1 (18)N2—C3—N4117.41 (10)
Li1—O6—H4A121 (2)C2—C3—N4129.07 (10)
C1—N1—N2—C30.25 (13)N1—C1—C2—C30.92 (14)
Li1—N1—N2—C3169.92 (12)N3—C1—C2—C3175.50 (12)
C1—N1—N2—Li1i176.01 (12)N1—C1—C2—C2ii174.43 (9)
Li1—N1—N2—Li1i5.8 (2)N3—C1—C2—C2ii9.15 (19)
Li1—O1—N3—O2173.04 (11)N1—N2—C3—C20.36 (14)
Li1—O1—N3—C14.91 (15)Li1i—N2—C3—C2175.68 (11)
Li1i—O3—N4—O4175.74 (11)N1—N2—C3—N4179.10 (11)
Li1i—O3—N4—C34.81 (14)Li1i—N2—C3—N43.07 (17)
N2—N1—C1—C20.78 (14)C1—C2—C3—N20.76 (14)
Li1—N1—C1—C2169.73 (11)C2ii—C2—C3—N2174.58 (10)
N2—N1—C1—N3175.99 (10)C1—C2—C3—N4179.32 (13)
Li1—N1—C1—N313.49 (18)C2ii—C2—C3—N44.0 (2)
O2—N3—C1—N1166.52 (12)O3—N4—C3—N22.02 (17)
O1—N3—C1—N111.53 (18)O4—N4—C3—N2178.50 (11)
O2—N3—C1—C29.7 (2)O3—N4—C3—C2179.46 (12)
O1—N3—C1—C2172.21 (13)O4—N4—C3—C20.0 (2)
Symmetry codes: (i) x+1, y+1, z; (ii) x+1/2, y+1/2, z.
Hydrogen-bond geometry (Å, º) top
D—H···AD—HH···AD···AD—H···A
O5—H1···O4iii0.87 (2)2.02 (2)2.8818 (14)174 (2)
O5—H2A···O2iv0.87 (2)2.13 (2)3.0010 (15)177 (4)
O5—H2B···O6v0.87 (2)2.02 (2)2.8555 (17)163 (3)
O6—H3···O1vi0.82 (3)2.24 (3)2.9995 (17)153 (2)
O6—H4A···O6vii0.92 (4)2.01 (4)2.905 (2)163 (4)
O6—H4B···O5viii0.95 (5)1.93 (5)2.8555 (17)164 (4)
Symmetry codes: (iii) x+1/2, y, z+1/2; (iv) x+1, y+1/2, z+1/2; (v) x+3/2, y, z+1/2; (vi) x, y+1/2, z1/2; (vii) x+3/2, y+1/2, z; (viii) x+3/2, y, z1/2.
 

Funding information

This work was supported by the Ministry of Education and Science of Ukraine (Project No. 22BF037–11) and the National Research Foundation of Ukraine (Project No. 2020.02/0071).

References

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 citationBrandenburg, K. (1999). DIAMOND. Crystal Impact GbR, Bonn, Germany.  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.  CSD CrossRef IUCr Journals Google Scholar
First citationDrukenmüller, I. E., Klapötke, T. M., Morgenstern, Y., Rusan, M. & Stierstorfer, J. (2014). Z. Anorg. Allg. Chem. 640, 2139–2148.  Google Scholar
First citationDufter, A. M. W., Klapötke, T. M., Rusan, M., Schweiger, A. & Stierstorfer, J. (2020). ChemPlusChem, 85, 2044–2050.  CSD CrossRef CAS PubMed Google Scholar
First citationDufter-Münster, A. M. W., Harter, A. G., Klapötke, T. M., Reinhardt, E., Römer, J. & Stierstorfer, J. (2022). Eur. J. Inorg. Chem. e202101048.  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 citationFarrugia, L. J. (2012). J. Appl. Cryst. 45, 849–854.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationGlück, J., Klapötke, T. M., Rusan, M., Sabatini, J. J. & Stierstorfer, J. (2017). Angew. Chem. Int. Ed. 56, 16507–16509.  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 citationGural'skiy, I. A., Escudero, D., Frontera, A., Solntsev, P. V., Rusanov, E. B., Chernega, A. N., Krautscheid, H. & Domasevitch, K. V. (2009). Dalton Trans. pp. 2856–2864.  Google Scholar
First citationHirshfeld, F. L. (1977). Theor. Chim. Acta, 44, 129–138.  CrossRef CAS Web of Science 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 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 citationOlsher, U., Izatt, R. M., Bradshaw, J. S. & Dalley, N. K. (1991). Chem. Rev. 91, 137–164.  CrossRef CAS Web of Science Google Scholar
First citationPonomarova, V. V., Komarchuk, V. V., Boldog, I., Krautscheid, H. & Domasevitch, K. V. (2013). CrystEngComm, 15, 8280–8287.  Web of Science CSD CrossRef CAS Google Scholar
First citationRavi, P. (2015). J. Mol. Struct. 1079, 433–447.  CrossRef CAS Google Scholar
First citationRossi, M., Rickles, L. F. & Glusker, J. P. (1983). Acta Cryst. C39, 987–990.  CSD CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationSabatini, J. J. (2018). Prop. Explos. Pyrotech. 43, 28–37.  CrossRef CAS Google Scholar
First citationShannon, R. D. (1976). Acta Cryst. A32, 751–767.  CrossRef CAS IUCr Journals Web of Science 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 citationSpackman, M. A. & McKinnon, J. J. (2002). CrystEngComm, 4, 378–392.  Web of Science CrossRef CAS Google Scholar
First citationStoe & Cie (2000). IPDS Software. Stoe & Cie GmbH, Darmstadt, Germany.  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. https://crystalexplorer.scb.uwa.edu.au/Google Scholar
First citationVeluthaparambath, R. V. P., Krishna, V., Pancharatna, P. D. & Saha, B. K. (2023). Cryst. Growth Des. 23, 442–449.  CrossRef CAS Google Scholar

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

Journal logoCRYSTALLOGRAPHIC
COMMUNICATIONS
ISSN: 2056-9890
Follow Acta Cryst. E
Sign up for e-alerts
Follow Acta Cryst. on Twitter
Follow us on facebook
Sign up for RSS feeds