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

Journal logoCRYSTALLOGRAPHIC
COMMUNICATIONS
ISSN: 2056-9890

Crystal structure and Hirshfeld surface analysis of 4,4′-(propane-1,3-diyl)bis­(4H-1,2,4-triazol-1-ium) penta­fluorido­oxidovanadate(V)

CROSSMARK_Color_square_no_text.svg

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: senchyk.ganna@gmail.com

Edited by W. T. A. Harrison, University of Aberdeen, Scotland (Received 21 April 2020; accepted 28 April 2020; online 1 May 2020)

In the structure of the title salt, (C7H12N6)[VOF5], second-order Jahn–Teller distortion of the coordination octa­hedra around V ions is reflected by coexistence of short V—O bonds [1.5767 (12) Å] and trans-positioned long V—F bonds [2.0981 (9) Å], with four equatorial V—F distances being inter­mediate in magnitude [1.7977 (9)–1.8913 (9) Å]. Hydrogen bonding of the anions is restricted to F-atom acceptors only, with particularly strong N–H⋯F inter­actions [N⋯F = 2.5072 (15) Å] established by axial and cis-positioned equatorial F atoms. Hirshfeld surface analysis indicates that the most important inter­actions are overwhelmingly H⋯F/F⋯H, accounting for 74.4 and 36.8% of the contacts for the individual anions and cations, respectively. Weak CH⋯F and CH⋯N bonds are essential for generation of three-dimensional structure.

1. Chemical context

Significant second-order Jahn–Teller distortions are inherent to the coordination octa­hedra of the [MvOF5]2− series (Mv = V, Nb, Ta) of ions (Ok et al., 2006[Ok, K. M., Halasyamani, P. S., Casanova, D., Llunell, M., Alemany, P. & Alvarez, S. (2006). Chem. Mater. 18, 3176-3183.]; Welk et al., 2002[Welk, M. E., Norquist, A. J., Arnold, F. P., Stern, C. L. & Poeppelmeier, K. R. (2002). Inorg. Chem. 41, 5119-5125.]). The resulting polar symmetry of the anions could be exploited as the origin of bulk polarity when imprinted on the structures of non-centrosymmetric coordination and hydrogen bonded solids (Halasyamani, 2010[Halasyamani, P. S. (2010). Functional Oxides, edited by D. W. Bruce, D. O'Hare & R. I. Walton, pp. 1-39. London: Wiley.]). Such supra­molecular synthesis with oxofluoride building blocks extends existing approaches for the development of non-centrosymmetric crystals, which attract significant inter­est for electro-optical applications (Gautier & Poeppelmeier, 2013[Gautier, R. & Poeppelmeier, K. R. (2013). Cryst. Growth Des. 13, 4084-4091.]).

One can anti­cipate that [VOF5]2− systems will show this effect to a particular extent since the vanadium ions experience a much larger out-of-centre displacement towards an apical O-ligand compared with their Nb and Ta analogues (Ok et al., 2006[Ok, K. M., Halasyamani, P. S., Casanova, D., Llunell, M., Alemany, P. & Alvarez, S. (2006). Chem. Mater. 18, 3176-3183.]). This feature generates a larger dipole moment as well as mitigating against orientational disorder of the anions in crystal structures (Sharko et al., 2018[Sharko, A. V., Erhart, O., Krautscheid, H. & Domasevitch, K. V. (2018). Acta Cryst. C74, 929-935.]). However, the supra­molecular behaviour of the [VOF5]2− anions is less predictable and it is strikingly different from that of the most extensively examined Nb and Ta systems. Welk et al. (2000[Welk, M. E., Norquist, A. J. P., Stern, C. L. & Poeppelmeier, K. R. (2000). Inorg. Chem. 39, 3946-3947.]) noted the very weak O-coordinating ability of the [VOF5]2− anions serving as F-donor ligands only but the hydrogen-bond acceptor ability of the O atoms is less addressed. Distal inter­actions of the C—H⋯O type are relevant to the structure of (H2bipy)[VOF5] (bipy is 4,4′-bi­pyridine; Gautier et al., 2015[Gautier, R., Gautier, R., Chang, K. B. & Poeppelmeier, K. R. (2015). Inorg. Chem. 54, 1712-1719.]), but surprisingly, no hydrogen bonding at all was observed for the O atoms in (H2En)[VOF5] (En is ethyl­enedi­amine; Rieskamp & Mattes, 1976[Rieskamp, H. & Mattes, R. (1976). Z. Naturforsch. Teil B, 31, 1453-1455.]). In addition, the possible competitiveness of the O atoms with respect to other weak hydrogen-bond acceptors does not appear to have been considered so far.

[Scheme 1]

With this in mind, we now describe the synthesis and structure of the title salt, (C7H12N62+)·[VOF5]2−, which gives insight into the hydrogen-bonding behaviour of [VOF5]2− anions when combined with the bitopic nitro­gen-rich 4,4′-(propane-1,3-diyl)bis(4H-1,2,4-triazol-1-ium) dication. This cation provides different kinds of hydrogen-bond donor sites complemented by triazole-N acceptors, which are relevant to many types of coordination and hydrogen-bonded systems (Senchyk et al., 2017[Senchyk, G. A., Lysenko, A. B., Domasevitch, K. V., Erhart, O., Henfling, S., Krautscheid, H., Rusanov, E. B., Krämer, K. W., Decurtins, S. & Liu, S.-X. (2017). Inorg. Chem. 56, 12952-12966.]; Lysenko et al., 2010[Lysenko, A. B., Senchyk, G. A., Lincke, J., Lässig, D., Fokin, A. A., Butova, E. D., Schreiner, P. R., Krautscheid, H. & Domasevitch, K. V. (2010). Dalton Trans. 39, 4223-4231.]).

2. Structural commentary

The mol­ecular structure of the title compounds is shown in Fig. 1[link]. The distorted coordination octa­hedra around the V ions comprise very short V1—O1 bonds of 1.5767 (12) Å and long bonds with trans-positioned F1 ligands [V1—F1 = 2.0981 (9) Å], which define the local polar axis of the anion. Four equatorial V—F bonds [mean 1.8295 (9) Å, Table 1[link]] are inter­mediate in length. That the anion geometry is sensitive to the hydrogen-bond environment is evidenced by the elongation of the V1—F4 bonds [1.8913 (9) Å], with the F4 atoms involved in a strong N—H⋯F inter­action (Table 2[link]). The central ion deviates from the centroid of its six ligand atoms by d = 0.242 Å towards the O-vertex. This is reminiscent of the geometrical features of the [VOF5]2− anions in the salts with (H2bipy)2+ (d = 0.268 Å; Gautier et al., 2015[Gautier, R., Gautier, R., Chang, K. B. & Poeppelmeier, K. R. (2015). Inorg. Chem. 54, 1712-1719.]) and (H2En)2+ cations (d = 0.272 Å; Rieskamp & Mattes, 1976[Rieskamp, H. & Mattes, R. (1976). Z. Naturforsch. Teil B, 31, 1453-1455.]).

Table 1
Selected geometric parameters (Å, °)

V1—O1 1.5767 (12) V1—F3 1.8228 (10)
V1—F5 1.7977 (9) V1—F4 1.8913 (9)
V1—F2 1.8062 (9) V1—F1 2.0981 (9)
       
O1—V1—F5 97.49 (6) F2—V1—F4 165.81 (4)
O1—V1—F2 97.75 (6) F3—V1—F4 85.92 (4)
F5—V1—F2 91.86 (5) O1—V1—F1 179.08 (6)
O1—V1—F3 96.57 (6) C1—N1—N2 111.79 (11)
F5—V1—F3 164.75 (5) C2—N2—N1 103.46 (12)
F2—V1—F3 92.06 (5) C3—N4—N5 111.17 (12)
O1—V1—F4 96.44 (6) C4—N5—N4 104.11 (12)
F5—V1—F4 86.68 (4)    
       
N3—C5—C6—C7 −63.73 (17) C5—C6—C7—N6 −171.58 (12)

Table 2
Hydrogen-bond geometry (Å, °)

D—H⋯A D—H H⋯A DA D—H⋯A
N1—H1N⋯F4i 0.87 1.76 2.6007 (14) 163
N4—H2N⋯F1ii 0.87 1.64 2.5072 (15) 173
C1—H1⋯F2 0.94 2.37 3.0962 (18) 133
C1—H1⋯N5iii 0.94 2.59 3.3122 (19) 134
C2—H2⋯F1iv 0.94 2.24 3.0163 (16) 139
C3—H3⋯F1v 0.94 2.42 3.2565 (18) 148
C3—H3⋯F5v 0.94 2.18 2.9980 (17) 144
C4—H4⋯N2vi 0.94 2.47 3.343 (2) 154
C5—H5A⋯F2 0.98 2.32 3.2039 (18) 150
C5—H5B⋯F1iv 0.98 2.54 3.2422 (18) 128
C6—H6A⋯F5v 0.98 2.50 3.4021 (19) 153
C7—H7A⋯F3 0.98 2.47 3.2728 (19) 139
C7—H7B⋯F4v 0.98 2.54 3.3522 (19) 141
Symmetry codes: (i) [-x+{\script{1\over 2}}, y-{\script{1\over 2}}, -z+{\script{1\over 2}}]; (ii) [-x+{\script{3\over 2}}, y+{\script{1\over 2}}, -z+{\script{1\over 2}}]; (iii) [-x+{\script{3\over 2}}, y-{\script{1\over 2}}, -z+{\script{1\over 2}}]; (iv) [x-{\script{1\over 2}}, -y+{\script{1\over 2}}, z-{\script{1\over 2}}]; (v) [x+{\script{1\over 2}}, -y+{\script{1\over 2}}, z-{\script{1\over 2}}]; (vi) [x+{\script{1\over 2}}, -y+{\script{1\over 2}}, z+{\script{1\over 2}}].
[Figure 1]
Figure 1
The mol­ecular structure of the title compound with displacement ellipsoids drawn at the 50% probability level. Dotted lines indicate weak C—H⋯F hydrogen bonding.

The main geometrical parameters of the organic cations are very similar to those of the parent 1,3-propyl­enebitriazole ligand in complexes with metal ions (Senchyk et al., 2017[Senchyk, G. A., Lysenko, A. B., Domasevitch, K. V., Erhart, O., Henfling, S., Krautscheid, H., Rusanov, E. B., Krämer, K. W., Decurtins, S. & Liu, S.-X. (2017). Inorg. Chem. 56, 12952-12966.]). The dicationic structure, as the result of protonation of the N1 and N4 sites, is best reflected by differentiation of the angles involving the N atoms in the two triazolium rings: C—N(H)—N = 111.17 (12) and 111.79 (11)° versus C—N—N(H) = 103.46 (12) and 104.11 (12)° (Table 1[link]). A similar effect is known for the isoelectronic neutral pyrazole ring (Gospodinov et al., 2020[Gospodinov, I., Domasevitch, K. V., Unger, C. C., Klapötke, T. M. & Stierstorfer, J. (2020). Cryst. Growth Des. 20, 755-764.]). The protonation also results in a certain shortening of the N—N bonds [1.362 (2) Å], as may be compared with N—N = 1.3918 (15) Å for the neutral and non-coordinated triazole rings in the adamantane derivative (Lysenko et al., 2019[Lysenko, A. B., Senchyk, G. A., Domasevitch, K. V., Henfling, S., Erhart, O., Krautscheid, H., Neves, P., Valente, A., Pillinger, M. & Gonçalves, I. S. (2019). Inorg. Chem. 58, 16424-16433.]). The methyl­ene linkage adopts a trans–gauche conformation with the corresponding torsion angles C5—C6—C7—N6 of −171.58 (12)° and N3—C5—C6—C7 of −63.73 (17)°. A diversity of metal complexes suggest nearlys equal occurrence of trans–gauche and all-trans sequences for the present moiety (Senchyk et al., 2017[Senchyk, G. A., Lysenko, A. B., Domasevitch, K. V., Erhart, O., Henfling, S., Krautscheid, H., Rusanov, E. B., Krämer, K. W., Decurtins, S. & Liu, S.-X. (2017). Inorg. Chem. 56, 12952-12966.]).

3. Supra­molecular features

The three-dimensional packing of the title compound is mediated by hydrogen bonding and two kinds of stacking inter­actions. Two strong N—H⋯F hydrogen bonds employ the most underbonded axial F1 atoms of the anion and the cis-positioned F4 atom (Fig. 2[link]). Thus the primary pattern exactly follows the cis-directing preferences of the [VOF5]2− anions, as suggested by Poeppelmeier and co-workers (Welk et al., 2000[Welk, M. E., Norquist, A. J. P., Stern, C. L. & Poeppelmeier, K. R. (2000). Inorg. Chem. 39, 3946-3947.]; Gautier et al., 2015[Gautier, R., Gautier, R., Chang, K. B. & Poeppelmeier, K. R. (2015). Inorg. Chem. 54, 1712-1719.]). More distal inter­actions are relevant to weaker CH donors (Table 2[link]). In total they support nine C—H⋯F contacts with a cut-off-limit of H⋯F = 2.56 Å, which is the sum of the van der Waals radii of these species (Rowland & Taylor, 1996[Rowland, R. S. & Taylor, R. (1996). J. Phys. Chem. 100, 7384-7391.]). The role of the triazole CH groups is notable: in addition to the shortest contacts with the F acceptors [H⋯F = 2.18–2.42 Å], they also form weak C—H⋯N bonds with triazole-N atoms [H⋯N = 2.47 and 2.59 Å; C⋯N = 3.3122 (19) and 3.343 (2) Å]. There are no N/C—H⋯O bonds at all and the shortest H⋯O contact of 2.84 Å considerably exceeds the sum of the corresponding van der Waals radii (2.68 Å; Rowland & Taylor, 1996[Rowland, R. S. & Taylor, R. (1996). J. Phys. Chem. 100, 7384-7391.]). It should be stressed that even such a weak acceptor as the N atom of the cationic moiety is a preferable site for hydrogen bonding, instead of the O atom of the [VOF5]2− anion. For the aliphatic portion of the structure, C—H⋯F inter­actions are longer and presumably weaker, whereas shorter H⋯F contacts [2.32 Å] correspond to the triazole-linked methyl­ene groups, as these are more polarized and acidic.

[Figure 2]
Figure 2
(a) Fragment of the double hydrogen-bonded chain showing the cis-directing function of the [VOF5]2− anions (with respect to the strongest N—H⋯F hydrogen-bond donors) and short anion⋯π contact subtended by the triazole ring to the O1/F2/F5 face of the anion. (b) Structure of the hydrogen-bonded layer, viewed nearly down the c axis, with the strongest hydrogen bonds and two kinds of stacking inter­actions indicated by blue and red dotted lines, respectively. [Symmetry codes: (i) −x + [{1\over 2}], y − [{1\over 2}], −z + [{1\over 2}]; (ii) −x + [{3\over 2}], y + [{1\over 2}], −z + [{1\over 2}]; (iv) [{3\over 2}] − x, −[{1\over 2}] + y, [{1\over 2}] − z; (vii) −x + [{1\over 2}], y + [{1\over 2}], −z + [{1\over 2}]; (viii) −x, −y, −z.]

Primary strong N—H⋯F bonding links the ionic counterparts into chains, which aggregate forming layers parallel to the ab plane. In a complement to the weak C—H⋯F bonds, these layers are sustained by two types of stacks (Fig. 2[link]). The first of these may be regarded as an inter­action between the triazolium ring to the F2/F5/O1 face of the anion, with an inter­planar angle of 12.60 (9)° and centroid-to-centroid distance of 3.064 (2) Å. This inter­action is favourable, as a kind of recently recognized anion⋯π bonding (Bauzá et al., 2016[Bauzá, A., Mooibroek, T. J. & Frontera, A. (2016). CrystEngComm, 18, 10-23.]) and it is responsible for the generation of a very short contact: F5⋯C3i = 2.7296 (15) Å [symmetry code: (i) −x + [{1\over 2}], y − [{1\over 2}], −z + [{1\over 2}]]. The second type may concern the stacking of the inversion–related triazolium rings. However, a relatively large inter­centroid distance of 3.626 (2) Å and slippage angle of 64.2 (2)° indicate a lack of overlap (Janiak, 2000[Janiak, C. (2000). J. Chem. Soc. Dalton Trans. pp. 3885-3896.]). Taking into account also the zero contribution of C⋯C contacts to the Hirshfeld surface of the cation (see below), one may postulate rather the ion–dipole inter­action of two triazolium N—NH+ sites, with the N1⋯N2viii separation of 3.2926 (18) Å [symmetry code: (viii) −x, −y, −z].

The packing of the layers extends the structure in the third dimension. For every next layer of the succession, the direction of the primary N—H⋯F bonded chains is inclined by 56.8° to the direction of chains from the preceding layer (Fig. 3[link]). Links between the layers represent most of the weak inter­actions, such as C—H⋯N bonds and C—H⋯F bonds with the aliphatic CH donors.

[Figure 3]
Figure 3
(a) Projection of the structure on the bc plane showing the extensive C—H⋯F and C—H⋯N inter­actions. A single hydrogen-bonded chain is marked red. (b) View down the c axis showing the inclined orientation of the hydrogen-bonded chains sustaining adjacent layers. Two separate layers are indicated in blue and red. [Symmetry code: (vi) x + [{1\over 2}], −y + [{1\over 2}], z + [{1\over 2}].]

4. Hirshfeld analysis

Supra­molecular inter­actions in the title structure were further accessed and visualized 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 cation mapped over dnorm using a fixed colour scale of −0.8385 (red) to 1.3445 (blue) a.u. indicates a number of red spots related to hydrogen-bond contacts. Particularly prominent spots are associated with the strongest N—H⋯F bonds. However, even the C—H⋯F inter­action with the weakest of the present donors (central CH2 group of the tri­methyl­ene linkage) is reflected by a red spot on the surface (Fig. 4[link]). The contribution of different kinds of inter­atomic contacts to the Hirshfeld surfaces of the individual cations and anions is shown in the fingerprint plots of Figs. 5[link] and 6[link]. Hydrogen-bond inter­actions (H⋯F and H⋯N/N⋯H) account for more than 60% of the contacts of the cations. The strong N—H⋯F bonding is reflected as a very sharp feature pointing to the lower left of the plot, with a shortest contact of 1.5 Å. The more distal H⋯N/N⋯H contacts (22.9%) are identified by a pair of shorter and diffuse spikes. There is no indication of directional H⋯O bonding: the plot represents a rather diffuse collection of points between the above features, with the shortest contact at 2.8 Å. A minor contribution of H⋯O contacts (5.3%) to the entire surface originates in the triazolium/F2,F5,O1 stack, but not in the C—H⋯O inter­actions.

[Figure 4]
Figure 4
The Hirshfeld surface of the cation mapped over dnorm in the colour range −0.8385 (red) to 1.3445 (blue) a.u., in the environment of the N—H⋯F and C—H⋯F hydrogen-bonded anions. [Symmetry codes: (i) −x + [{1\over 2}], y − [{1\over 2}], −z + [{1\over 2}]; (ii) −x + [{3\over 2}], y + [{1\over 2}], −z + [{1\over 2}]; (iv) x − [{1\over 2}], −y + [{1\over 2}], z − [{1\over 2}].]
[Figure 5]
Figure 5
Two-dimensional fingerprint plots for the cations of the title compound, and delineated into the principal contributions of H⋯F, H⋯N/N⋯H, H⋯O and C⋯F contacts. Other important contacts are H⋯H (18.5%), H⋯C/C⋯H (3.4%) and N⋯N (3.0%).
[Figure 6]
Figure 6
Two-dimensional fingerprint plots for the [VOF5]2− anions, showing the very different character of the F⋯H and O⋯H contacts. Very short F⋯C contacts are also readily detectable. Other important contacts are F⋯C(N) and O⋯C(N) contributing 6.8 and 5.4%, respectively.

The same conclusion may be reached when considering the surface area of the [VOF5]2− anions. The inter­molecular contacts in this case are overwhelmingly of the type F⋯H (74.4%). In addition to this very sharp spike, the plot clearly reveals the more subtle feature of anion⋯π bonding, which appears as a short spike at 2.7 Å (Fig. 6[link]). In total, the corres­ponding F⋯C(N) and O⋯C(N) contacts account for 12.2% of the anion contacts. There are no close C⋯C contacts, while the contribution of N⋯N contacts (3.0%) is perceptible in the fingerprint plots of the cations. As noted above, this indicates pairwise ion–dipole inter­actions of the N—NH+ fragments, with a lack of genuine ππ inter­actions.

5. Synthesis and crystallization

The bitriazole was prepared in a yield of 33% by the acid-catalysed condensation of 1,3-di­amino­propane and N,N-di­methyl­formamide azine (Lysenko et al., 2010[Lysenko, A. B., Senchyk, G. A., Lincke, J., Lässig, D., Fokin, A. A., Butova, E. D., Schreiner, P. R., Krautscheid, H. & Domasevitch, K. V. (2010). Dalton Trans. 39, 4223-4231.]). To prepare the title compound, the bitriazole (71.2 mg, 0.40 mmol), V2O5 (18.2 mg, 0.10 mmol), 0.84 ml of 7% aqueous HF solution (3.0 mmol) and 2 ml of water were placed in a Teflon vessel and heated in a steel bomb at 413 K for 24 h. Cooling to room temperature over a period of 48 h afforded colourless crystals of the title salt, in a yield of 27 mg (40%). Analysis (%) calculated for C7H12F5N6OV: C 24.57, H 3.54, N 24.57; found: C 24.38, H 3.49, N 24.70.

6. Refinement

Crystal data, data collection and structure refinement details are summarized in Table 3[link]. All hydrogen atoms were located and then refined as riding with N—H = 0.87 Å, C—H (triazole) = 0.94 Å and C—H (CH2) = 0.98 Å; Uiso(H) = 1.2Ueq(CH) and 1.5Ueq(NH).

Table 3
Experimental details

Crystal data
Chemical formula (C7H12N6)[VOF5]
Mr 342.17
Crystal system, space group Monoclinic, P21/n
Temperature (K) 213
a, b, c (Å) 6.5915 (4), 12.1969 (10), 15.5669 (10)
β (°) 97.617 (8)
V3) 1240.47 (15)
Z 4
Radiation type Mo Kα
μ (mm−1) 0.87
Crystal size (mm) 0.25 × 0.22 × 0.20
 
Data collection
Diffractometer Stoe IPDS
Absorption correction Numerical [X-RED (Stoe & Cie, 2001[Stoe & Cie (2001). X-RED. Stoe & Cie GmbH, Darmstadt, Germany.]) and X-SHAPE (Stoe & Cie, 1999[Stoe & Cie (1999). X-SHAPE. Stoe & Cie GmbH, Darmstadt, Germany.])]
Tmin, Tmax 0.272, 0.303
No. of measured, independent and observed [I > 2σ(I)] reflections 10733, 2965, 2513
Rint 0.028
(sin θ/λ)max−1) 0.663
 
Refinement
R[F2 > 2σ(F2)], wR(F2), S 0.028, 0.080, 1.01
No. of reflections 2965
No. of parameters 181
H-atom treatment H-atom parameters constrained
Δρmax, Δρmin (e Å−3) 0.37, −0.28
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.]), SHELXL2018/1 (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: SHELXL2018/1 (Sheldrick, 2015); molecular graphics: Diamond (Brandenburg, 1999); software used to prepare material for publication: WinGX (Farrugia, 2012).

4,4'-(Propane-1,3-diyl)bis(4H-1,2,4-triazol-1-ium) pentafluoridooxidovanadate(V) top
Crystal data top
(C7H12N6)[VF5O]F(000) = 688
Mr = 342.17Dx = 1.832 Mg m3
Monoclinic, P21/nMo Kα radiation, λ = 0.71073 Å
a = 6.5915 (4) ÅCell parameters from 8000 reflections
b = 12.1969 (10) Åθ = 3.2–28.1°
c = 15.5669 (10) ŵ = 0.87 mm1
β = 97.617 (8)°T = 213 K
V = 1240.47 (15) Å3Prism, colorless
Z = 40.25 × 0.22 × 0.20 mm
Data collection top
Stoe Image plate diffraction system
diffractometer
2513 reflections with I > 2σ(I)
Radiation source: fine-focus sealed tubeRint = 0.028
φ oscillation scansθmax = 28.1°, θmin = 3.2°
Absorption correction: numerical
[X-RED (Stoe & Cie, 2001) and X-SHAPE (Stoe & Cie, 1999)]
h = 78
Tmin = 0.272, Tmax = 0.303k = 1616
10733 measured reflectionsl = 1919
2965 independent 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)] = 0.028Hydrogen site location: difference Fourier map
wR(F2) = 0.080H-atom parameters constrained
S = 1.01 w = 1/[σ2(Fo2) + (0.0603P)2]
where P = (Fo2 + 2Fc2)/3
2965 reflections(Δ/σ)max < 0.001
181 parametersΔρmax = 0.37 e Å3
0 restraintsΔρmin = 0.28 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*/Ueq
V10.13004 (4)0.20736 (2)0.37557 (2)0.02234 (9)
O10.0838 (2)0.25118 (10)0.33158 (9)0.0429 (3)
F10.41341 (13)0.14981 (7)0.43607 (6)0.0320 (2)
F20.18077 (15)0.11480 (7)0.29038 (6)0.0361 (2)
F30.28439 (17)0.31538 (8)0.33574 (7)0.0422 (2)
F40.14112 (15)0.29233 (6)0.47771 (6)0.0310 (2)
F50.03593 (14)0.09762 (7)0.43649 (6)0.0329 (2)
N10.27600 (18)0.00076 (9)0.03807 (8)0.0250 (2)
H1N0.3177830.0683860.0432510.038*
N20.2150 (2)0.04943 (10)0.03920 (8)0.0290 (3)
N30.19669 (16)0.16124 (9)0.07027 (8)0.0214 (2)
N40.88170 (17)0.52627 (10)0.14823 (8)0.0260 (3)
H2N0.9615820.5666650.1211210.039*
N50.8618 (2)0.53656 (11)0.23383 (9)0.0333 (3)
N60.66746 (17)0.40329 (9)0.17358 (8)0.0237 (2)
C10.2642 (2)0.06477 (11)0.10324 (10)0.0262 (3)
H10.2969380.0474890.1622960.031*
C20.1676 (2)0.14804 (11)0.01688 (10)0.0283 (3)
H20.1191400.2033420.0564120.034*
C30.7655 (2)0.44809 (10)0.11223 (9)0.0238 (3)
H30.7528340.4269190.0537040.029*
C40.7297 (2)0.46124 (12)0.24712 (10)0.0312 (3)
H40.6829930.4484550.3007370.037*
C50.1521 (2)0.26078 (11)0.11809 (10)0.0261 (3)
H5A0.1399850.2409490.1782000.031*
H5B0.0202640.2908960.0921530.031*
C60.3158 (2)0.34849 (11)0.11766 (10)0.0270 (3)
H6A0.3303790.3672370.0575610.032*
H6B0.2723020.4146990.1458440.032*
C70.5216 (2)0.31175 (11)0.16388 (10)0.0274 (3)
H7A0.5041790.2829230.2212180.033*
H7B0.5757710.2527130.1307940.033*
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
V10.03279 (14)0.01556 (13)0.01820 (14)0.00023 (8)0.00168 (9)0.00033 (7)
O10.0483 (7)0.0392 (6)0.0376 (7)0.0135 (5)0.0074 (5)0.0005 (5)
F10.0311 (4)0.0312 (4)0.0318 (5)0.0036 (3)0.0028 (4)0.0077 (3)
F20.0530 (5)0.0307 (4)0.0234 (5)0.0052 (4)0.0009 (4)0.0081 (3)
F30.0674 (7)0.0296 (4)0.0319 (6)0.0133 (4)0.0154 (5)0.0046 (4)
F40.0505 (5)0.0198 (4)0.0241 (5)0.0032 (3)0.0097 (4)0.0041 (3)
F50.0405 (5)0.0260 (4)0.0308 (5)0.0120 (3)0.0003 (4)0.0041 (3)
N10.0291 (6)0.0180 (5)0.0274 (7)0.0027 (4)0.0017 (4)0.0023 (4)
N20.0369 (6)0.0267 (6)0.0231 (7)0.0030 (5)0.0026 (5)0.0002 (5)
N30.0229 (5)0.0179 (5)0.0229 (6)0.0006 (4)0.0015 (4)0.0021 (4)
N40.0254 (5)0.0254 (5)0.0260 (7)0.0005 (4)0.0005 (4)0.0025 (5)
N50.0408 (7)0.0314 (6)0.0256 (7)0.0024 (5)0.0034 (5)0.0038 (5)
N60.0277 (5)0.0211 (5)0.0217 (6)0.0023 (4)0.0009 (4)0.0009 (4)
C10.0306 (7)0.0219 (6)0.0248 (8)0.0025 (5)0.0012 (5)0.0043 (5)
C20.0376 (7)0.0240 (6)0.0226 (8)0.0046 (5)0.0018 (6)0.0046 (5)
C30.0260 (6)0.0222 (6)0.0223 (7)0.0036 (5)0.0005 (5)0.0009 (5)
C40.0409 (8)0.0315 (7)0.0205 (8)0.0012 (6)0.0020 (6)0.0014 (6)
C50.0300 (7)0.0210 (6)0.0279 (8)0.0020 (5)0.0057 (5)0.0026 (5)
C60.0329 (7)0.0177 (6)0.0291 (8)0.0017 (5)0.0008 (5)0.0010 (5)
C70.0303 (7)0.0195 (6)0.0319 (8)0.0005 (5)0.0016 (6)0.0035 (5)
Geometric parameters (Å, º) top
V1—O11.5767 (12)N5—C41.301 (2)
V1—F51.7977 (9)N6—C31.3387 (19)
V1—F21.8062 (9)N6—C41.3615 (19)
V1—F31.8228 (10)N6—C71.4683 (17)
V1—F41.8913 (9)C1—H10.9400
V1—F12.0981 (9)C2—H20.9400
N1—C11.3019 (19)C3—H30.9400
N1—N21.3621 (18)C4—H40.9400
N1—H1N0.8700C5—C61.5201 (19)
N2—C21.3014 (19)C5—H5A0.9800
N3—C11.3361 (16)C5—H5B0.9800
N3—C21.354 (2)C6—C71.5160 (19)
N3—C51.4740 (17)C6—H6A0.9800
N4—C31.3022 (17)C6—H6B0.9800
N4—N51.362 (2)C7—H7A0.9800
N4—H2N0.8700C7—H7B0.9800
O1—V1—F597.49 (6)N1—C1—H1126.5
O1—V1—F297.75 (6)N3—C1—H1126.5
F5—V1—F291.86 (5)N2—C2—N3111.73 (13)
O1—V1—F396.57 (6)N2—C2—H2124.1
F5—V1—F3164.75 (5)N3—C2—H2124.1
F2—V1—F392.06 (5)N4—C3—N6107.69 (13)
O1—V1—F496.44 (6)N4—C3—H3126.2
F5—V1—F486.68 (4)N6—C3—H3126.2
F2—V1—F4165.81 (4)N5—C4—N6111.47 (14)
F3—V1—F485.92 (4)N5—C4—H4124.3
O1—V1—F1179.08 (6)N6—C4—H4124.3
F5—V1—F182.13 (4)N3—C5—C6112.85 (12)
F2—V1—F183.11 (4)N3—C5—H5A109.0
F3—V1—F183.73 (5)C6—C5—H5A109.0
F4—V1—F182.71 (4)N3—C5—H5B109.0
C1—N1—N2111.79 (11)C6—C5—H5B109.0
C1—N1—H1N124.1H5A—C5—H5B107.8
N2—N1—H1N124.1C7—C6—C5112.37 (11)
C2—N2—N1103.46 (12)C7—C6—H6A109.1
C1—N3—C2106.01 (12)C5—C6—H6A109.1
C1—N3—C5127.58 (13)C7—C6—H6B109.1
C2—N3—C5126.34 (12)C5—C6—H6B109.1
C3—N4—N5111.17 (12)H6A—C6—H6B107.9
C3—N4—H2N124.4N6—C7—C6110.88 (11)
N5—N4—H2N124.4N6—C7—H7A109.5
C4—N5—N4104.11 (12)C6—C7—H7A109.5
C3—N6—C4105.54 (12)N6—C7—H7B109.5
C3—N6—C7127.50 (12)C6—C7—H7B109.5
C4—N6—C7126.96 (13)H7A—C7—H7B108.1
N1—C1—N3107.01 (12)
C1—N1—N2—C20.38 (16)C7—N6—C3—N4179.39 (12)
C3—N4—N5—C40.12 (16)N4—N5—C4—N60.68 (17)
N2—N1—C1—N30.72 (16)C3—N6—C4—N51.21 (16)
C2—N3—C1—N10.74 (15)C7—N6—C4—N5179.41 (12)
C5—N3—C1—N1177.92 (12)C1—N3—C5—C6104.60 (16)
N1—N2—C2—N30.11 (16)C2—N3—C5—C678.76 (17)
C1—N3—C2—N20.54 (17)N3—C5—C6—C763.73 (17)
C5—N3—C2—N2177.76 (13)C3—N6—C7—C675.14 (17)
N5—N4—C3—N60.88 (15)C4—N6—C7—C6104.11 (16)
C4—N6—C3—N41.23 (14)C5—C6—C7—N6171.58 (12)
Hydrogen-bond geometry (Å, º) top
D—H···AD—HH···AD···AD—H···A
N1—H1N···F4i0.871.762.6007 (14)163
N4—H2N···F1ii0.871.642.5072 (15)173
C1—H1···F20.942.373.0962 (18)133
C1—H1···N5iii0.942.593.3122 (19)134
C2—H2···F1iv0.942.243.0163 (16)139
C3—H3···F1v0.942.423.2565 (18)148
C3—H3···F5v0.942.182.9980 (17)144
C4—H4···N2vi0.942.473.343 (2)154
C5—H5A···F20.982.323.2039 (18)150
C5—H5B···F1iv0.982.543.2422 (18)128
C6—H6A···F5v0.982.503.4021 (19)153
C7—H7A···F30.982.473.2728 (19)139
C7—H7B···F4v0.982.543.3522 (19)141
Symmetry codes: (i) x+1/2, y1/2, z+1/2; (ii) x+3/2, y+1/2, z+1/2; (iii) x+3/2, y1/2, z+1/2; (iv) x1/2, y+1/2, z1/2; (v) x+1/2, y+1/2, z1/2; (vi) x+1/2, y+1/2, z+1/2.
 

Funding information

The authors gratefully acknowledge support from the Deutsche Forschungsgemeinschaft, grant KR1675/12–1 (HK and KVD). This work was supported by the Ministry of Education and Science of Ukraine (project No. 19BF037–05).

References

First citationBauzá, A., Mooibroek, T. J. & Frontera, A. (2016). CrystEngComm, 18, 10–23.  Google Scholar
First citationBrandenburg, K. (1999). DIAMOND. Crystal Impact GbR, Bonn, Germany.  Google Scholar
First citationFarrugia, L. J. (2012). J. Appl. Cryst. 45, 849–854.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationGautier, R., Gautier, R., Chang, K. B. & Poeppelmeier, K. R. (2015). Inorg. Chem. 54, 1712–1719.  Web of Science CSD CrossRef CAS PubMed Google Scholar
First citationGautier, R. & Poeppelmeier, K. R. (2013). Cryst. Growth Des. 13, 4084–4091.  Web of Science CSD CrossRef CAS 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 citationHalasyamani, P. S. (2010). Functional Oxides, edited by D. W. Bruce, D. O'Hare & R. I. Walton, pp. 1–39. London: Wiley.  Google Scholar
First citationHirshfeld, F. L. (1977). Theor. Chim. Acta, 44, 129–138.  CrossRef CAS Web of Science Google Scholar
First citationJaniak, C. (2000). J. Chem. Soc. Dalton Trans. pp. 3885–3896.  Web of Science CrossRef Google Scholar
First citationLysenko, A. B., Senchyk, G. A., Domasevitch, K. V., Henfling, S., Erhart, O., Krautscheid, H., Neves, P., Valente, A., Pillinger, M. & Gonçalves, I. S. (2019). Inorg. Chem. 58, 16424–16433.  Web of Science CSD CrossRef CAS PubMed Google Scholar
First citationLysenko, A. B., Senchyk, G. A., Lincke, J., Lässig, D., Fokin, A. A., Butova, E. D., Schreiner, P. R., Krautscheid, H. & Domasevitch, K. V. (2010). Dalton Trans. 39, 4223–4231.  Web of Science CSD CrossRef CAS PubMed 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 citationOk, K. M., Halasyamani, P. S., Casanova, D., Llunell, M., Alemany, P. & Alvarez, S. (2006). Chem. Mater. 18, 3176–3183.  Web of Science CrossRef CAS Google Scholar
First citationRieskamp, H. & Mattes, R. (1976). Z. Naturforsch. Teil B, 31, 1453–1455.  CrossRef Google Scholar
First citationRowland, R. S. & Taylor, R. (1996). J. Phys. Chem. 100, 7384–7391.  CrossRef CAS Web of Science Google Scholar
First citationSenchyk, G. A., Lysenko, A. B., Domasevitch, K. V., Erhart, O., Henfling, S., Krautscheid, H., Rusanov, E. B., Krämer, K. W., Decurtins, S. & Liu, S.-X. (2017). Inorg. Chem. 56, 12952–12966.  Web of Science CSD CrossRef CAS PubMed Google Scholar
First citationSharko, A. V., Erhart, O., Krautscheid, H. & Domasevitch, K. V. (2018). Acta Cryst. C74, 929–935.  Web of Science CSD CrossRef IUCr Journals 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 (1999). X-SHAPE. Stoe & Cie GmbH, Darmstadt, Germany.  Google Scholar
First citationStoe & Cie (2000). IPDS Software. Stoe & Cie GmbH, Darmstadt, Germany.  Google Scholar
First citationStoe & Cie (2001). X-RED. 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 citationWelk, M. E., Norquist, A. J., Arnold, F. P., Stern, C. L. & Poeppelmeier, K. R. (2002). Inorg. Chem. 41, 5119–5125.  Web of Science CSD CrossRef PubMed CAS Google Scholar
First citationWelk, M. E., Norquist, A. J. P., Stern, C. L. & Poeppelmeier, K. R. (2000). Inorg. Chem. 39, 3946–3947.  Web of Science CSD CrossRef PubMed 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