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 5-hy­dr­oxy­penta­nehydrazide

crossmark logo

aDepartamento de Química, Universidade Federal do Paraná, Centro, Politécnico, Jardim das Américas, 81530-900, Curitiba-PR, Brazil
*Correspondence e-mail: marcelodoca@ufpr.br

Edited by L. Van Meervelt, Katholieke Universiteit Leuven, Belgium (Received 19 February 2024; accepted 1 April 2024; online 9 April 2024)

Carb­oxy­hydrazides are widely used in medicinal chemistry because of their medicinal properties and many drugs have been developed containing this functional group. A suitable inter­mediate to obtain potential hydrazide drug candidates is the title compound 5-hy­droxy­penta­nehydrazide, C5H12N2O2 (1). The aliphatic compound can react both via the hydroxyl and hydrazide moieties forming derivatives, which can inhibit Mycobacterium tuberculosis catalase-peroxidase (KatG) and consequently causes death of the pathogen. In this work, the hydrazide was obtained via a reaction of a lactone with hydrazine hydrate. The colourless prismatic single crystals belong to the ortho­rhom­bic space group Pca21. Regarding supra­molecular inter­actions, the compound shows classic medium to strong inter­molecular hydrogen bonds involving the hydroxyl and hydrazide groups. Besides, the three-dimensional packing also shows weak H⋯H and C⋯H contacts, as investigated by Hirshfeld surface analysis (HS) and fingerprint plots (FP).

1. Chemical context

Carboxyhydrazides are non-alkaline compounds that can be identified as hydrazines containing an acyl group as one of their substituents, thus they are of general formula R1–NR2–NR3R4, where R­1 is an acyl group and R2R4 are typically hydrogen atoms or alkyl substituents. These compounds, particularly those in which R2R4 are hydrogen atoms, present themselves as valuable functional groups for drug design, since compounds with this functional group and its derivatives tend to have biological activity as anti­depressants or anti­biotics, for example (Narang et al., 2012[Narang, R., Narasimhan, B. & Sharma, S. (2012). Curr. Med. Chem. 19, 569-612.]). The medicinal potentiality led to the development of several drugs containing this functional group, such as isoniazid, furazolidone, and isocarboxazide (Gegia et al., 2017[Gegia, M., Winters, N., Benedetti, A., van Soolingen, D. & Menzies, D. (2017). Lancet Infect. Dis. 17, 223-234.]).

[Scheme 1]

Hydrazides are usually formed via the reaction of hydrazine (usually obtained from its hydro­chloride or its hydrate) with acyl derivatives such as esters, acyl halides, or anhydrides (Huang et al., 2016[Huang, Y., Fang, G. & Liu, L. (2016). Natl. Sci. 3, 107-116.]). For example, lactones (cyclic esters) promptly react with hydrazine hydrate in a polar solvent such as methanol. In this work, δ-valerolactone was added to hydrazine to afford ortho­rhom­bic crystals of 5-hy­droxy­penta­nehydrazide 1 (Huang et al., 2016[Huang, Y., Fang, G. & Liu, L. (2016). Natl. Sci. 3, 107-116.]).

Compound 1 was first synthesized by Karakhanov and collaborators in 1969 (Karakhanov et al., 1969[Karakhanov, R., Shekhtman, N. & Zefirov, N. (1969). Chem. Heterocycl. Compd 5, 18.]). However, this is the first report describing the crystallographic features of 5-hy­droxy­penta­nehydrazide.

2. Structural commentary

The mol­ecule of 5-hy­droxy­penta­nehydrazide (Fig. 1[link]), crystallizes in the ortho­rhom­bic space group Pca21. The asymmetric unit comprises a unique mol­ecule of 1 with no atoms in special positions, as well as no solvent of crystallization. The C—N, C=O, and N—N bond lengths within the hydrazide group of 1.3376 (17) Å, 1.2375 (16) Å, and 1.4193 (14), respectively, are in agreement with the values reported for aliphatic compounds containing a hydrazide unit (Jensen, 1956[Jensen, L. (1956). J. Am. Chem. Soc. 78, 3993-3999.]; Lo et al., 2020[Lo, K. M., Lee, S. M. & Tiekink, E. R. (2020). Z. Krist. New Cryst. Struct. 235, 1257-1258.]; Kolesnikova et al., 2022[Kolesnikova, I., Rykov, A. & Shishkov, I. (2022). J. Mol. Struct. 1250, 139-447.]). Moreover, the short and unbranched carbon chain formed by atoms C1, C2, C3, C4, and C5, is located in a plane (blue plane in Fig. 2[link]) that makes an angle α of 54.8 (9)° relative to the plane containing the hydrazide atoms N1, N2, C1, C2 and O1 (grey plane in Fig. 2[link]).

[Figure 1]
Figure 1
View of the mol­ecular structure of 5-hy­droxy­penta­nehydrazide (1) with the atom-numbering scheme. Displacement ellipsoids are drawn at the 50% probability level.
[Figure 2]
Figure 2
Representation of the dihedral angle α formed by the planes containing carbon chain atoms C1, C2, C3, C4, and C5 (blue plane) and the hydrazine group atoms N1, N2, O1, C1, and C2 (grey plane) in 1. Grey: carbon; red: oxygen; blue: nitro­gen. H atoms atoms were omitted for clarity. Displacement ellipsoids are drawn at the 50% probability level.

The aforementioned conformational features, linked to the orientation of the hydrazide group, are a relatively common characteristic in compounds containing these groups linked to carbon chains. The α angle of 54.8 (9)° observed in compound 1 is consistent with values reported in the literature, regardless of the carbon chain size. Noteworthy, values include 54.54° for a compound with a twelve-carbon chain (Jensen, 1956[Jensen, L. (1956). J. Am. Chem. Soc. 78, 3993-3999.]), 56.33° for a nine-carbon chain (Jensen & Lingafelter, 1961[Jensen, L. H. & Lingafelter, E. C. (1961). Acta Cryst. 14, 507-520.]), and 57.08° for a six-carbon chain (Lee et al., 2016[Lee, S. M., Lo, K. M., Tan, S. L. & Tiekink, E. R. T. (2016). Acta Cryst. E72, 1390-1395.]).

3. Supra­molecular features

The three-dimensional packing of 1 (Fig. 3[link]) is characterized by several inter­molecular O–H⋯N and N–H⋯O hydrogen bonds involving the hydroxyl and hydrazine as hydrogen bond donor and acceptor groups (Table 1[link]). Among them, the strongest ones, regarding shortest H⋯acceptor distances and most linear donor—H⋯acceptor angles, are the N1—H3⋯O1ii [2.03 (2) Å; 166 (2)°; symmetry code: (ii) x, y + 1, z] and O2—H4⋯N2i [1.92 (3) Å; 172 (2)°; symmetry code: (i) −x + 1, −y + 2, z + [{1\over 2}]], involving the hydrazine and hydroxyl as donor group, respectively, and carbonyl and hydrazine moieties as acceptor groups, respectively. Besides, the hydrazine moiety also promotes hydrogen bonds of medium-force: N2—H1⋯O2iii [2.10 (2) Å; 147.2 (18)°; symmetry code: (iii) −x + [{3\over 2}], y, z − [{1\over 2}]], N2—H2⋯O1iv [2.59 (2) Å; 144.0 (16)°; symmetry code: (iv) x − [{1\over 2}], −y + 1, z] and N2—H2⋯O2v [2.57 (2) Å; 122.0 (16)°; symmetry code: (v) −x + 1, 1 − y, z − [{1\over 2}]]. The latter promotes the formation of chains with a C(9) graph-set motif running in the c-axis direction. A weak hydrogen bond of type C—H⋯O is also present in the crystal packing. Notably, thanks to this weak inter­action, the carbonyl group is the acceptor of a bifurcated hydrogen bond, sharing its electronic density with the C2—H2 and N2—H2 groups and resulting in a six-membered ring with an R21(6) graph-set motif in the b-axis direction (Fig. 4[link]a). The formation of seven-membered rings with an R23(7) graph-set motif is observed involving the H1–N2–N1–C1–O1⋯H2vi⋯O2iii moiety in the a-axis direction [symmetry code: (iii) −x + [{3\over 2}], y, z − [{1\over 2}]; (vi) x + [{1\over 2}], −y + 1, z; Fig. 4[link]b). Together, these inter­actions act cooperatively for the stability of 1 in the solid state (Sutor et al., 1962[Sutor, D. (1962). Nature, 195, 68-69.]; Domagała & Grabowski, 2005[Domagała, M. & Grabowski, S. J. (2005). J. Phys. Chem. A, 109, 5683-5688.]).

Table 1
Hydrogen-bond geometry (Å, °)

D—H⋯A D—H H⋯A DA D—H⋯A
O2—H4⋯N2i 0.87 (3) 1.92 (3) 2.7865 (15) 172 (2)
N1—H3⋯O1ii 0.85 (2) 2.03 (2) 2.8662 (14) 166 (2)
N2—H1⋯O2iii 0.91 (2) 2.10 (2) 2.9068 (16) 147.2 (18)
N2—H2⋯O1iv 0.82 (2) 2.59 (2) 3.2858 (16) 144.0 (16)
N2—H2⋯O2v 0.82 (2) 2.57 (2) 3.0789 (14) 122.0 (16)
C2—H2A⋯O1ii 0.99 2.57 3.4309 (15) 145
Symmetry codes: (i) [-x+1, -y+2, z+{\script{1\over 2}}]; (ii) [x, y+1, z]; (iii) [-x+{\script{3\over 2}}, y, z-{\script{1\over 2}}]; (iv) [x-{\script{1\over 2}}, -y+1, z]; (v) [-x+1, -y+1, z-{\script{1\over 2}}].
[Figure 3]
Figure 3
Crystal packing of 1 viewed along the b axis. Inter­molecular O—H⋯N, N—H⋯O and C—H⋯O hydrogen bonds are shown as dashed blue lines. H atoms not involved in hydrogen bonding were omitted for clarity.
[Figure 4]
Figure 4
(a) Representation of the six-membered ring formed by N—H⋯O and C–H⋯O hydrogen bonds (blue dashed lines) between adjacent mol­ecules in the b-axis direction. (b) Representation of the seven-membered ring formed by N—H⋯O hydrogen bonds (blue dashed lines) between carbonyl and hydrazide groups of adjacent mol­ecules in the a-axis direction. Symmetry codes: (i) x, −y + 1, z; (iii) −x + [{3\over 2}], y, z − 1; (vi) x + [{1\over 2}], −y + 1, z. Displacement ellipsoids are drawn at the 50% probability level.

The non-covalent inter­actions responsible for the crystal packing were also investigated by a Hirshfeld surface analysis (HS; Hirshfeld, 1977[Hirshfeld, F. L. (1977). Theor. Chim. Acta, 44, 129-138.]), performed with CrystalExplorer 21.5 (Spackman et al., 2021[Spackman, P. R., Turner, M. J., McKinnon, J. J., Wolff, S. K., Grimwood, D. J., Jayatilaka, D. & Spackman, M. A. (2021). J. Appl. Cryst. 54, 1006-1011.]). The Hirshfeld surface provides a three-dimensional representation that elucidates mol­ecular inter­actions through the mathematical distance functions di, denoting the distance from the surface to the nearest atom within it, and de, denoting the distance from the surface to the nearest atom outside of it. The normalization of the di and de distances by the van der Waals radius leads to the dnorm function, which enables the visualization of a surface that delineates regions involved in both accepting and donating inter­molecular inter­actions. A key component of this analysis entails the generation of 2D fingerprint plots (FP), providing two-dimensional representations of the Hirshfeld surface.

Using the dnorm function, expressed by a colour scale, this method describes the strength of inter­atomic inter­actions. Red and blue indicate inter­atomic contacts where the distance between atoms is smaller or larger, respectively, than the sum of the van der Waals radii of the atoms involved, while white indicates contacts with distances close to the sum of the van der Waals radii.

In the case of compound 1, the red colour in Fig. 5[link]a highlights the region of most intense contacts involving the nitro­gen, oxygen and hydrogen atoms from the hydrazide and hydroxyl groups with adjacent oxygen atoms. Fig. 5[link]b illustrates the nearest mol­ecules within the crystal packing, delineating the spatial arrangement of the shortest inter­actions. Meanwhile, blue surfaces, which indicate longer-range inter­actions, arise mainly from H⋯H contributions.

[Figure 5]
Figure 5
Representation of (a) the Hirshfeld surface for 1 plotted over dnorm and (b) illustration of the N—H⋯O, O—H⋯N and C—H⋯O inter­actions depicted by dashed green lines.

Fingerprint plots (FP) were generated to qu­antify the contribution of each inter­atomic inter­action to the supra­molecular structure. For this purpose, the di (x axis) and de (y axis) distances, expressed in Ångstroms, of the HS are used. For 1, the percentages of the surface area correspond to 64.7% for H⋯H, 26.2% for O⋯H/H⋯O, 7.5% for N⋯H/H⋯N, 1.2% for C⋯H/H⋯C, 0.3% for O⋯C/C⋯O and 0.1% for O⋯O inter­actions, as shown in Fig. 6[link].

[Figure 6]
Figure 6
Fingerprint plots for 1 showing the total contribution of individual inter­actions and those delineated into H⋯H, N⋯H/H⋯N, O⋯H/H⋯O, C⋯H/H⋯C, O⋯C/C⋯O and O⋯O inter­actions.

4. Database survey

A survey of the Cambridge Structural Database (CSD2023.2.0, version 5.45, November 2023; Groom et al., 2016[Groom, C. R., Bruno, I. J., Lightfoot, M. P. & Ward, S. C. (2016). Acta Cryst. B72, 171-179.]) revealed several similar structures. 5-Hy­droxy­penta­nehydrazide was first synthesized as a byproduct of the reaction of di­hydro­pyran and phenyl azide (Karakhanov et al., 1969[Karakhanov, R., Shekhtman, N. & Zefirov, N. (1969). Chem. Heterocycl. Compd 5, 18.]) and has never had its structural properties discussed, although it was first obtained in its crystalline form. The synthesis and crystallographic characterization of other similar aliphatic hydrazide derivatives has been reported: t-butyl hydrazine­carboxyl­ate (CSD refcode RENZUJ; Aitken & Slawin, 2022[Aitken, R. & Slawin, A. (2022). Molbank, M1482.]), α-cyano­acetohydrazide (CYACHZ; Chieh, 1973[Chieh, P. (1973). J. Chem. Soc. Perkin Trans. 2, pp. 1825-1828.]), n-dodeca­noic acid hydrazide (DDEAHN; Jensen, 1956[Jensen, L. (1956). J. Am. Chem. Soc. 78, 3993-3999.]), hexa­nedihydrazide (MUYRIK; Lo et al., 2020[Lo, K. M., Lee, S. M. & Tiekink, E. R. (2020). Z. Krist. New Cryst. Struct. 235, 1257-1258.]), n-nona­noic acid hydrazide (NONACH; Jensen & Lingafelter, 1961[Jensen, L. H. & Lingafelter, E. C. (1961). Acta Cryst. 14, 507-520.]) and n-octa­noic acid hydrazide (ZZZOMM; Jensen & Lingafelter, 1953[Jensen, L. H. & Lingafelter, E. C. (1953). Acta Cryst. 6, 300-301.]).

5. Synthesis and crystallization

To a round-bottom flask, δ-valerolactone (100 mg, 9.99 mmol), hydrazine hydrate (200 mg, 4.0 mmol) and 5 mL of methanol were added. The resulting solution was maintained stirring under reflux conditions for 24 h. The solution was then allowed to cool slowly to room temperature. After 20 minutes, a solid started to precipitate in the flask. The solid, which was filtered off and air dried, afforded 115.0 mg of colourless crystals of 5-hy­droxy­penta­nehydrazide (1) in 88.0% yield. The melting point (375–379 K) was in accordance with literature (Karakhanov et al., 1969[Karakhanov, R., Shekhtman, N. & Zefirov, N. (1969). Chem. Heterocycl. Compd 5, 18.]).

6. Refinement

Crystal data, data collection and structure refinement details are summarized in Table 2[link]. The hydrogen atoms of the carbon chain were included in idealized positions with C—H distances set to 0.99 Å and refined using a riding model with Uiso(H) = 1.2Ueq(C); the other hydrogen atoms were located in difference-Fourier maps and were refined freely.

Table 2
Experimental details

Crystal data
Chemical formula C5H12N2O2
Mr 132.17
Crystal system, space group Orthorhombic, Pca21
Temperature (K) 100
a, b, c (Å) 7.1686 (5), 4.8491 (3), 19.1276 (14)
V3) 664.90 (8)
Z 4
Radiation type Mo Kα
μ (mm−1) 0.10
Crystal size (mm) 0.32 × 0.16 × 0.13
 
Data collection
Diffractometer Bruker D8 Venture/Photon 100 CMOS
Absorption correction Multi-scan (SADABS; Krause et al., 2015[Krause, L., Herbst-Irmer, R., Sheldrick, G. M. & Stalke, D. (2015). J. Appl. Cryst. 48, 3-10.])
Tmin, Tmax 0.731, 0.746
No. of measured, independent and observed [I > 2σ(I)] reflections 26421, 1567, 1527
Rint 0.029
(sin θ/λ)max−1) 0.655
 
Refinement
R[F2 > 2σ(F2)], wR(F2), S 0.022, 0.059, 1.08
No. of reflections 1567
No. of parameters 98
No. of restraints 1
H-atom treatment H atoms treated by a mixture of independent and constrained refinement
Δρmax, Δρmin (e Å−3) 0.21, −0.19
Absolute structure Flack x determined using 732 quotients [(I+)−(I)]/[(I+)+(I)] (Parsons et al., 2013[Parsons, S., Flack, H. D. & Wagner, T. (2013). Acta Cryst. B69, 249-259.])
Absolute structure parameter −0.1 (2)
Computer programs: APEX4 (Bruker, 2022[Bruker (2022). APEX4. Bruker AXS Inc., Madison, Wisconsin, USA.]), SAINT (Bruker, 2019[Bruker (2019). SAINT. Bruker AXS Inc., Madison, Wisconsin, USA.]), SHELXT2015 (Sheldrick, 2015a[Sheldrick, G. M. (2015a). Acta Cryst. A71, 3-8.]), SHELXL2019/2 (Sheldrick, 2015b[Sheldrick, G. M. (2015b). Acta Cryst. C71, 3-8.]), DIAMOND (Brandenburg & Putz, 1999[Brandenburg, K. & Putz, H. (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

5-Hydroxypentanehydrazide top
Crystal data top
C5H12N2O2Dx = 1.320 Mg m3
Mr = 132.17Mo Kα radiation, λ = 0.71073 Å
Orthorhombic, Pca21Cell parameters from 9932 reflections
a = 7.1686 (5) Åθ = 4.2–28.7°
b = 4.8491 (3) ŵ = 0.10 mm1
c = 19.1276 (14) ÅT = 100 K
V = 664.90 (8) Å3Prism, colourless
Z = 40.32 × 0.16 × 0.13 mm
F(000) = 288
Data collection top
Bruker D8 Venture/Photon 100 CMOS
diffractometer
1567 independent reflections
Radiation source: fine-focus sealed tube1527 reflections with I > 2σ(I)
Detector resolution: 10.4167 pixels mm-1Rint = 0.029
φ and ω scansθmax = 27.8°, θmin = 4.3°
Absorption correction: multi-scan
(SADABS; Krause et al., 2015)
h = 99
Tmin = 0.731, Tmax = 0.746k = 66
26421 measured reflectionsl = 2525
Refinement top
Refinement on F2Secondary atom site location: difference Fourier map
Least-squares matrix: fullHydrogen site location: mixed
R[F2 > 2σ(F2)] = 0.022H atoms treated by a mixture of independent and constrained refinement
wR(F2) = 0.059 w = 1/[σ2(Fo2) + (0.0379P)2 + 0.0769P]
where P = (Fo2 + 2Fc2)/3
S = 1.08(Δ/σ)max < 0.001
1567 reflectionsΔρmax = 0.21 e Å3
98 parametersΔρmin = 0.19 e Å3
1 restraintAbsolute structure: Flack x determined using 732 quotients [(I+)-(I-)]/[(I+)+(I-)] (Parsons et al., 2013)
Primary atom site location: dualAbsolute structure parameter: 0.1 (2)
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
O10.58285 (13)0.44491 (18)0.38116 (6)0.0149 (2)
O20.71437 (14)0.7429 (2)0.71584 (5)0.0151 (2)
N10.51145 (15)0.8787 (2)0.34743 (5)0.0105 (2)
N20.42516 (17)0.7936 (2)0.28423 (6)0.0122 (2)
C10.58629 (18)0.6966 (3)0.39182 (6)0.0097 (2)
C30.58864 (17)0.6955 (2)0.52353 (6)0.0110 (2)
H3A0.4539990.7404430.5252390.013*
H3B0.6010040.4922510.5222130.013*
C40.68313 (18)0.8053 (3)0.58941 (6)0.0115 (3)
H4A0.8187470.7685360.5865160.014*
H4B0.6655521.0076240.5917650.014*
C50.60650 (19)0.6753 (3)0.65582 (7)0.0139 (3)
H5A0.4766480.7387690.6630340.017*
H5B0.6039200.4724470.6501580.017*
C20.67320 (18)0.8176 (2)0.45691 (6)0.0106 (2)
H2A0.6545421.0199190.4569260.013*
H2B0.8091320.7816230.4564170.013*
H10.514 (3)0.715 (4)0.2573 (12)0.024 (5)*
H20.345 (3)0.680 (4)0.2935 (10)0.016 (4)*
H30.517 (3)1.052 (4)0.3531 (11)0.020 (5)*
H40.664 (4)0.890 (5)0.7337 (13)0.034 (6)*
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
O10.0227 (5)0.0074 (4)0.0146 (4)0.0001 (3)0.0016 (4)0.0006 (3)
O20.0204 (5)0.0158 (4)0.0092 (4)0.0034 (4)0.0026 (4)0.0033 (4)
N10.0161 (5)0.0070 (5)0.0084 (5)0.0001 (4)0.0012 (4)0.0011 (4)
N20.0160 (5)0.0111 (5)0.0094 (5)0.0012 (4)0.0029 (4)0.0016 (4)
C10.0106 (5)0.0096 (5)0.0090 (6)0.0006 (4)0.0033 (4)0.0005 (4)
C30.0127 (5)0.0120 (5)0.0082 (5)0.0013 (4)0.0005 (4)0.0012 (5)
C40.0132 (6)0.0120 (6)0.0092 (5)0.0010 (4)0.0003 (5)0.0009 (5)
C50.0170 (6)0.0166 (6)0.0081 (5)0.0031 (5)0.0008 (5)0.0007 (5)
C20.0136 (6)0.0093 (5)0.0088 (5)0.0014 (5)0.0004 (5)0.0002 (4)
Geometric parameters (Å, º) top
O1—C11.2375 (16)C3—C21.5303 (16)
O2—C51.4223 (16)C3—H3A0.9900
O2—H40.87 (3)C3—H3B0.9900
N1—C11.3376 (17)C4—C51.5208 (17)
N1—N21.4193 (14)C4—H4A0.9900
N1—H30.85 (2)C4—H4B0.9900
N2—H10.91 (2)C5—H5A0.9900
N2—H20.82 (2)C5—H5B0.9900
C1—C21.5109 (16)C2—H2A0.9900
C3—C41.5265 (16)C2—H2B0.9900
C5—O2—H4106.4 (17)C5—C4—H4A109.1
C1—N1—N2121.55 (11)C3—C4—H4A109.1
C1—N1—H3123.8 (15)C5—C4—H4B109.1
N2—N1—H3114.6 (15)C3—C4—H4B109.1
N1—N2—H1107.3 (14)H4A—C4—H4B107.8
N1—N2—H2108.6 (13)O2—C5—C4112.48 (10)
H1—N2—H2109.6 (19)O2—C5—H5A109.1
O1—C1—N1122.60 (12)C4—C5—H5A109.1
O1—C1—C2121.81 (11)O2—C5—H5B109.1
N1—C1—C2115.58 (11)C4—C5—H5B109.1
C4—C3—C2112.13 (10)H5A—C5—H5B107.8
C4—C3—H3A109.2C1—C2—C3111.87 (10)
C2—C3—H3A109.2C1—C2—H2A109.2
C4—C3—H3B109.2C3—C2—H2A109.2
C2—C3—H3B109.2C1—C2—H2B109.2
H3A—C3—H3B107.9C3—C2—H2B109.2
C5—C4—C3112.62 (10)H2A—C2—H2B107.9
N2—N1—C1—O10.98 (19)O1—C1—C2—C355.33 (16)
N2—N1—C1—C2180.00 (11)N1—C1—C2—C3123.70 (11)
C2—C3—C4—C5177.29 (11)C4—C3—C2—C1176.88 (10)
C3—C4—C5—O2170.30 (10)
Hydrogen-bond geometry (Å, º) top
D—H···AD—HH···AD···AD—H···A
O2—H4···N2i0.87 (3)1.92 (3)2.7865 (15)172 (2)
N1—H3···O1ii0.85 (2)2.03 (2)2.8662 (14)166 (2)
N2—H1···O2iii0.91 (2)2.10 (2)2.9068 (16)147.2 (18)
N2—H2···O1iv0.82 (2)2.59 (2)3.2858 (16)144.0 (16)
N2—H2···O2v0.82 (2)2.57 (2)3.0789 (14)122.0 (16)
C2—H2A···O1ii0.992.573.4309 (15)145
Symmetry codes: (i) x+1, y+2, z+1/2; (ii) x, y+1, z; (iii) x+3/2, y, z1/2; (iv) x1/2, y+1, z; (v) x+1, y+1, z1/2.
 

Acknowledgements

The authors thank the Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq), Edital PROIND 2020/UFPR and Coordenacão de Aperfeiçoamento de Pessoal de Nível Superior (CAPES) for fellowships. The author's contributions are as follows. GAJ: methodology, writing – original draft; JSCN: formal analysis, resources; FSS: data curation, visualization; MLBR: validation; FGSP: investigation; PPD: writing, review and editing; MGMD: conceptual­ization, supervision.

Funding information

Funding for this research was provided by: Coordenação de Aperfeiçoamento de Pessoal de Nível Superior (grant No. 23038.003745/2021-31); Pró-Reitoria de Planejamento, Orçamentos e Finanças - UFPR - PROIND 2020 (grant No. 3527).

References

First citationAitken, R. & Slawin, A. (2022). Molbank, M1482.  Web of Science CSD CrossRef Google Scholar
First citationBrandenburg, K. & Putz, H. (1999). DIAMOND. Crystal Impact GbR, Bonn, Germany.  Google Scholar
First citationBruker (2019). SAINT. Bruker AXS Inc., Madison, Wisconsin, USA.  Google Scholar
First citationBruker (2022). APEX4. Bruker AXS Inc., Madison, Wisconsin, USA.  Google Scholar
First citationChieh, P. (1973). J. Chem. Soc. Perkin Trans. 2, pp. 1825–1828.  CSD CrossRef Web of Science Google Scholar
First citationDomagała, M. & Grabowski, S. J. (2005). J. Phys. Chem. A, 109, 5683–5688.  Web of Science PubMed Google Scholar
First citationFarrugia, L. J. (2012). J. Appl. Cryst. 45, 849–854.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationGegia, M., Winters, N., Benedetti, A., van Soolingen, D. & Menzies, D. (2017). Lancet Infect. Dis. 17, 223–234.  Web of Science CrossRef CAS PubMed 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 citationHuang, Y., Fang, G. & Liu, L. (2016). Natl. Sci. 3, 107–116.  CAS Google Scholar
First citationJensen, L. (1956). J. Am. Chem. Soc. 78, 3993–3999.  CSD CrossRef CAS Web of Science Google Scholar
First citationJensen, L. H. & Lingafelter, E. C. (1953). Acta Cryst. 6, 300–301.  CSD CrossRef IUCr Journals Web of Science Google Scholar
First citationJensen, L. H. & Lingafelter, E. C. (1961). Acta Cryst. 14, 507–520.  CSD CrossRef IUCr Journals Web of Science Google Scholar
First citationKarakhanov, R., Shekhtman, N. & Zefirov, N. (1969). Chem. Heterocycl. Compd 5, 18.  CrossRef Google Scholar
First citationKolesnikova, I., Rykov, A. & Shishkov, I. (2022). J. Mol. Struct. 1250, 139–447.  Web of Science CrossRef Google Scholar
First citationKrause, L., Herbst-Irmer, R., Sheldrick, G. M. & Stalke, D. (2015). J. Appl. Cryst. 48, 3–10.  Web of Science CSD CrossRef ICSD CAS IUCr Journals Google Scholar
First citationLee, S. M., Lo, K. M., Tan, S. L. & Tiekink, E. R. T. (2016). Acta Cryst. E72, 1390–1395.  CSD CrossRef IUCr Journals Google Scholar
First citationLo, K. M., Lee, S. M. & Tiekink, E. R. (2020). Z. Krist. New Cryst. Struct. 235, 1257–1258.  CAS Google Scholar
First citationNarang, R., Narasimhan, B. & Sharma, S. (2012). Curr. Med. Chem. 19, 569–612.  Web of Science CrossRef CAS PubMed Google Scholar
First citationParsons, S., Flack, H. D. & Wagner, T. (2013). Acta Cryst. B69, 249–259.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationSheldrick, G. M. (2015a). Acta Cryst. A71, 3–8.  Web of Science CrossRef IUCr Journals Google Scholar
First citationSheldrick, G. M. (2015b). Acta Cryst. C71, 3–8.  Web of Science CrossRef IUCr Journals Google Scholar
First citationSpackman, P. R., Turner, M. J., McKinnon, J. J., Wolff, S. K., Grimwood, D. J., Jayatilaka, D. & Spackman, M. A. (2021). J. Appl. Cryst. 54, 1006–1011.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationSutor, D. (1962). Nature, 195, 68–69.  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