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

Journal logoCRYSTALLOGRAPHIC
COMMUNICATIONS
ISSN: 2056-9890

Low-temperature crystal structure of 4-chloro-1H-pyrazole

crossmark logo

aDepartment of Chemistry and Biochemistry, 11200 SW 8th Street, Miami, FL 33199, USA
*Correspondence e-mail: rraptis@fiu.edu

Edited by M. Zeller, Purdue University, USA (Received 4 June 2021; accepted 17 August 2021; online 24 August 2021)

The structure of 4-chloro-1H-pyrazole, C3H3ClN2, at 170 K has ortho­rhom­bic (Pnma) symmetry and is isostructural to its bromo analogue. Data were collected at low temperature since 4-chloro-1H-pyrazole sublimes when subjected to the localized heat produced by X-rays. The structure displays inter­molecular N—H⋯N hydrogen bonding and the packing features a trimeric mol­ecular assembly bis­ected by a mirror plane (m normal to b) running through one chlorine atom, one carbon atom and one N—N bond. The asymmetric unit therefore consists of one and one-half 4-chloro-1H-pyrazole mol­ecules. Thus, the N—H proton is crystallographically disordered over two positions of 50% occupancy each.

1. Chemical context

Pyrazoles are a family of five-membered, π-excess aromatic heterocycles with adjacent nitro­gen atoms (Krishnakumar et al., 2011[Krishnakumar, V., Jayamani, N. & Mathammal, R. (2011). Spectrochim. Acta A Mol. Biomol. Spectrosc. 79, 1959-1968.]; Karrouchi et al., 2018[Karrouchi, K., Radi, S., Ramli, Y., Taoufik, J., Mabkhot, Y., Al-aizari, F. A. & Ansar, M. (2018). Molecules, 23, 134.]). They serve as scaffolds for numerous pharmaceutically active compounds and agrochemicals as they are stable and amenable to substitution (Naim et al., 2016[Naim, M. J., Alam, O., Nawaz, F., Alam, J. & Alam, P. (2016). J. Pharm. Bioallied Sci. 8, 2-17.]). They readily coordinate to metals, forming a large group of complexes where typically pyrazoles are monodentate or bridging bidentate pyrazolido ligands (Pettinari et al., 2010[Pettinari, C., Masciocchi, N., Pandolfo, L. & Pucci, D. (2010). Chem. Eur. J. 16, 1106-1123.]; Viciano-Chumillas et al., 2010[Viciano-Chumillas, M., Tanase, S., de Jongh, L. J. & Reedijk, J. (2010). Eur. J. Inorg. Chem. pp. 3403-3418.]). In the solid state, pyrazoles form hydrogen-bonded aggregates whose topology depends on the steric bulk of peripheral substituents and the acidity of the N1—H proton and basicity of the N2 atom. For example, 4-bromo-pyrazole and 3-methyl-pyrazole form approximately planar triangular assemblies (Foces-Foces et al., 1999[Foces-Foces, C., Llamas-Saiz, A. L. & Elguero, J. (1999). Z. Kristallogr. 214, 237-241.]; Goddard et al., 1999[Goddard, R., Claramunt, R. M., Escolástico, C. & Elguero, J. (1999). New J. Chem. 23, 237-240.]), 3,5-diphenyl-pyrazole and 3,5-bis­(tri­fluoro­meth­yl)pyrazole form tetra­nuclear assemblies (Raptis et al., 1993[Raptis, R. G., Staples, R. J., King, C. & Fackler, J. P. (1993). Acta Cryst. C49, 1716-1719.]; Alkorta et al., 1999[Alkorta, I., Elguero, J., Donnadieu, B., Etienne, M., Jaffart, J., Schagen, D. & Limbach, H.-H. (1999). New J. Chem. 23, 1231-1237.]), while pyrazole forms a one-dimensional polymeric chain (Krebs Larsen et al., 1970[Krebs Larsen, F., Lehmann, M. S., Søtofte, I. & Rasmussen, S. E. (1970). Acta Chem. Scand. 24, 3248-3258.]).

[Scheme 1]

2. Structural commentary

4-Chloro-1H-pyrazole crystallizes with one and one-half mol­ecules in the asymmetric unit (Z′ = 1.5) as shown in Fig. 1[link]. The second mol­ecule is bis­ected by a mirror plane normal to b (x, 3/4, z), which runs through C5, Cl2, and the N3—N3i bond [symmetry code: (i) x, [{1\over 2}] − y, z]. As a result of this mirror plane, the NH protons are crystallographically disordered over two positions. As with the bromo analogue (Foces-Foces et al., 1999[Foces-Foces, C., Llamas-Saiz, A. L. & Elguero, J. (1999). Z. Kristallogr. 214, 237-241.]), 4-chloro-1H-pyrazole crystallizes in the Pnma space group and forms trimeric units (Fig. 2[link]).

[Figure 1]
Figure 1
Perspective view of the asymmetric unit of 4-chloro-1H-pyrazole. Displacement ellipsoids are shown at 50% probability and half the disordered protons are removed for clarity.
[Figure 2]
Figure 2
Packing of 4-chloro-1H-pyrazole, viewed parallel to the c axis, showing the formation of trimeric units. Half the disordered protons have been removed for clarity.

Also disordered in this structure are the C and N atoms. Without disorder, pyrazoles have one `pyrrole-like' side and one `pyridine-like' side, as was discussed in the neutron diffraction study of 1H-pyrazole (Krebs Larsen et al., 1970[Krebs Larsen, F., Lehmann, M. S., Søtofte, I. & Rasmussen, S. E. (1970). Acta Chem. Scand. 24, 3248-3258.]). The C—N, C—C, and C—H bonds on either side of the mol­ecule are crystallographically distinct and resemble either pyrrole or pyridine. However, due to the disorder of the N—H protons in the current structure, only average positions of the C and N atoms have been obtained. Therefore, the C—N, C—C, and C—H bonds on either side of the mol­ecule are indistinguishable. This is most apparent in the C—N bonds. In the current structure, the C—N bonds are 1.335 (2), 1.334 (2), and 1.334 (2) Å for C1—N1, C3—N2, and C4 —N3, respectively. In the 1H-pyrazole structure, the C—N bond lengths are 1.356 and 1.350 Å for the `pyrrole-like' side and the `pyridine-like' side, respectively. The C—N bond lengths of the current structure are in agreement with those previously reported in the 4-bromo analogue with C—N bond lengths of 1.343 (10), 1.331 (10), and 1.327 (10) Å (Foces-Foces et al., 1999[Foces-Foces, C., Llamas-Saiz, A. L. & Elguero, J. (1999). Z. Kristallogr. 214, 237-241.]). Furthermore, these bond lengths are also in agreement with the 4-chloro-1H-pyrazol-2-ium chloride salt, which exhibits C—N bond lengths of 1.334 (2) and 1.331 (2) Å (Farmiloe et al., 2019[Farmiloe, S. E., Berdiell, I. C. & Halcrow, M. E. (2019). CSD Communication (deposition No. 1944671). CCDC, Cambridge, England. https://doi.org/10.5517/ccdc.csd.cc238lbb]). The N—N bonds of the present mol­ecule are 1.346 (2) and 1.345 (3) Å, for N1—N2 and N3—N3i, respectively, and are similar to those of 4-methyl-1H-pyrazole (which is refined without proton disorder in Pca21) with N—N bond lengths of 1.343 (2), 1.344 (2), and 1.349 (2) Å (Goddard et al., 1999[Goddard, R., Claramunt, R. M., Escolástico, C. & Elguero, J. (1999). New J. Chem. 23, 237-240.]). However, the N—N bonds of the 4-bromo analogue are slightly shorter at 1.335 (9) Å each (Foces-Foces et al., 1999[Foces-Foces, C., Llamas-Saiz, A. L. & Elguero, J. (1999). Z. Kristallogr. 214, 237-241.]).

3. Supra­molecular features

Pyrazoles are known to crystallize in a variety of hydrogen-bonded motifs – dimers, trimers, tetra­mers, and catemers (Infantes & Motherwell, 2004[Infantes, L. & Motherwell, S. (2004). Struct. Chem. 15, 173-184.]; Pérez & Riera, 2009[Pérez, J. & Riera, L. (2009). Eur. J. Inorg. Chem. pp. 4913-4925.]). The current structure of 4-chloro-1H-pyrazole crystallizes in trimeric units, as do the 4-bromo and 4-methyl analogues. The inter­molecular N1⋯N1i, N2⋯N3ii, and N3⋯N2ii [symmetry codes: (i) x, −y + [{3\over 2}], z; (ii) −x + 1, −y + 1, −z] are 2.885 (3), 2.858 (2), and 2.858 (2), respectively, as shown in Table 1[link]. These values are in agreement with other inter­molecular hydrogen-bond inter­actions of pyrazoles. Just as with the 4-bromo derivative, 4-chloro-1H-pyrazole packs in a herringbone arrangement when viewed down the b axis (Fig. 3[link]). The current structure exhibits no π-stacking inter­actions.

Table 1
Hydrogen-bond geometry (Å, °)

D—H⋯A D—H H⋯A DA D—H⋯A
N1—H1A⋯N1i 0.88 2.03 2.885 (3) 165
N2—H2⋯N3ii 0.88 1.99 2.8582 (19) 169
N3—H3A⋯N2ii 0.88 1.99 2.8582 (19) 169
Symmetry codes: (i) [x, -y+{\script{3\over 2}}, z]; (ii) [-x+1, -y+1, -z].
[Figure 3]
Figure 3
Packing of 4-chloro-1H-pyrazole, viewed parallel to the b axis, showing the formation of a herringbone motif.

4. Synthesis and crystallization

4-Chloro-1H-pyrazole was purchased commercially and crystals were grown from the slow evaporation of a methyl­ene chloride solution.

5. Refinement

Crystal data, data collection and structure refinement details are summarized in Table 2[link]. The data were collected at 170 K on a Bruker D8 CMOS diffractometer equipped with a Photon II detector. CrysAlis PRO was used for scaling and absorption correction. All C—H protons were refined freely while the N—H protons were fixed using an AFIX command and constrained to half occupancy due to the proton disorder.

Table 2
Experimental details

Crystal data
Chemical formula C3H3ClN2
Mr 102.52
Crystal system, space group Orthorhombic, Pnma
Temperature (K) 170
a, b, c (Å) 14.9122 (10), 17.6410 (9), 4.9878 (3)
V3) 1312.13 (14)
Z 12
Radiation type Mo Kα
μ (mm−1) 0.69
Crystal size (mm) 0.29 × 0.14 × 0.08
 
Data collection
Diffractometer Bruker D8 CMOS
Absorption correction Multi-scan (CrysAlis PRO; Rigaku OD, 2019[Rigaku OD (2019). CrysAlis PRO. Rigaku Oxford Diffraction, Yarnton, England.])
Tmin, Tmax 0.760, 1.000
No. of measured, independent and observed [I > 2σ(I)] reflections 25891, 1798, 1481
Rint 0.043
(sin θ/λ)max−1) 0.682
 
Refinement
R[F2 > 2σ(F2)], wR(F2), S 0.033, 0.092, 1.06
No. of reflections 1798
No. of parameters 97
H-atom treatment H atoms treated by a mixture of independent and constrained refinement
Δρmax, Δρmin (e Å−3) 0.27, −0.26
Computer programs: APEX3 (Brker, 2020[Bruker (2020). APEX3. Bruker AXS Inc., Madison, Wisconsin, USA.]), CrysAlis PRO (Rigaku OD, 2019[Rigaku OD (2019). CrysAlis PRO. Rigaku Oxford Diffraction, Yarnton, England.]), SHELXS (Sheldrick, 2008[Sheldrick, G. M. (2008). Acta Cryst. A64, 112-122.]), SHELXL2018/3 (Sheldrick, 2015[Sheldrick, G. M. (2015). Acta Cryst. C71, 3-8.]), and OLEX2 (Dolomanov et al., 2009[Dolomanov, O. V., Bourhis, L. J., Gildea, R. J., Howard, J. A. K. & Puschmann, H. (2009). J. Appl. Cryst. 42, 339-341.]).

To remove the proton disorder, an attempt was made to refine the mol­ecule in the non-centrosymmetric Pn21a space group, which is also consistent with the systematic absences. In Pn21a, hydrogen atoms were refined in both possible positions – on the odd-labeled N atoms or on the even-labeled N atoms. However, as Foces-Foces et al. (1999[Foces-Foces, C., Llamas-Saiz, A. L. & Elguero, J. (1999). Z. Kristallogr. 214, 237-241.]) found with the 4-bromo-analogue, refinement in the lower symmetry space group did not improve the refinement. For the trial Pn21a structure with protons on the even-labeled N atoms, the R1 and wR2 values slightly increased to 3.67% and 9.67%, respectively, and the shifting of the carbon atoms could not be consolidated. For the trial Pn21a structure with protons on the odd-labeled N atoms, the R1 and wR2 values increased even more to 3.71% and 9.69%, respectively, and the coordinates of the non-proton­ated N atoms could not converge. Therefore, the best refined model was chosen to be in the centrosymmetric Pnma space group with NH protons disordered over two positions.

Supporting information


Computing details top

Data collection: APEX3 (Bruker, 2020); cell refinement: CrysAlis PRO (Rigaku OD, 2019); data reduction: CrysAlis PRO (Rigaku OD, 2019); program(s) used to solve structure: SHELXS (Sheldrick, 2008); program(s) used to refine structure: SHELXL2018/3 (Sheldrick, 2015); molecular graphics: OLEX2 (Dolomanov et al., 2009); software used to prepare material for publication: OLEX2 (Dolomanov et al., 2009).

4-Chloro-1H-pyrazole top
Crystal data top
C3H3ClN2Dx = 1.557 Mg m3
Mr = 102.52Mo Kα radiation, λ = 0.71073 Å
Orthorhombic, PnmaCell parameters from 6701 reflections
a = 14.9122 (10) Åθ = 2.3–30.2°
b = 17.6410 (9) ŵ = 0.69 mm1
c = 4.9878 (3) ÅT = 170 K
V = 1312.13 (14) Å3Rect. Prism, clear colourless
Z = 120.28 × 0.14 × 0.08 mm
F(000) = 624
Data collection top
Bruker D8 CMOS
diffractometer
1481 reflections with I > 2σ(I)
ω scansRint = 0.043
Absorption correction: multi-scan
(CrysAlisPro; Rigaku OD, 2019)
θmax = 29.0°, θmin = 2.3°
Tmin = 0.760, Tmax = 1.000h = 2020
25891 measured reflectionsk = 2323
1798 independent reflectionsl = 66
Refinement top
Refinement on F2Primary atom site location: dual
Least-squares matrix: fullSecondary atom site location: difference Fourier map
R[F2 > 2σ(F2)] = 0.033Hydrogen site location: mixed
wR(F2) = 0.092H atoms treated by a mixture of independent and constrained refinement
S = 1.06 w = 1/[σ2(Fo2) + (0.0418P)2 + 0.5547P]
where P = (Fo2 + 2Fc2)/3
1798 reflections(Δ/σ)max = 0.001
97 parametersΔρmax = 0.27 e Å3
0 restraintsΔρmin = 0.26 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)
Cl10.65912 (3)0.45830 (2)0.73672 (9)0.04510 (15)
C10.66638 (11)0.61504 (9)0.6893 (4)0.0375 (3)
C20.63595 (10)0.54554 (8)0.6008 (3)0.0330 (3)
N10.63427 (9)0.66822 (7)0.5248 (3)0.0390 (3)
H1A0.6441410.7172420.5388440.047*0.5
C30.58468 (11)0.56001 (10)0.3776 (3)0.0389 (4)
N20.58435 (9)0.63462 (8)0.3340 (3)0.0405 (3)
H20.5562360.6579390.2026490.049*0.5
Cl20.70256 (4)0.2500000.59928 (11)0.03892 (16)
C50.62343 (15)0.2500000.3495 (4)0.0326 (4)
N30.52730 (9)0.28813 (7)0.0432 (3)0.0395 (3)
H3A0.4944320.3172370.0610950.047*0.5
C40.58543 (12)0.31226 (10)0.2276 (3)0.0396 (4)
H30.5571 (14)0.5266 (11)0.253 (4)0.043 (5)*
H10.7031 (12)0.6279 (11)0.839 (4)0.040 (5)*
H40.5962 (12)0.3661 (12)0.261 (4)0.051 (6)*
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
Cl10.0525 (3)0.0303 (2)0.0525 (3)0.00392 (16)0.00952 (19)0.00342 (16)
C10.0379 (8)0.0335 (8)0.0412 (8)0.0017 (6)0.0027 (7)0.0020 (7)
C20.0340 (7)0.0281 (7)0.0371 (8)0.0020 (6)0.0008 (6)0.0006 (6)
N10.0408 (7)0.0330 (6)0.0431 (7)0.0007 (6)0.0017 (6)0.0012 (6)
C30.0429 (9)0.0345 (8)0.0391 (8)0.0028 (6)0.0053 (7)0.0032 (7)
N20.0450 (8)0.0383 (7)0.0384 (7)0.0053 (6)0.0022 (6)0.0028 (6)
Cl20.0416 (3)0.0395 (3)0.0356 (3)0.0000.0053 (2)0.000
C50.0359 (11)0.0311 (10)0.0309 (10)0.0000.0011 (9)0.000
N30.0432 (7)0.0370 (7)0.0384 (7)0.0068 (6)0.0036 (6)0.0001 (6)
C40.0483 (10)0.0306 (8)0.0401 (8)0.0052 (7)0.0035 (7)0.0040 (7)
Geometric parameters (Å, º) top
Cl1—C21.7169 (15)N2—H20.8800
C1—C21.380 (2)Cl2—C51.716 (2)
C1—N11.335 (2)C5—C41.377 (2)
C1—H10.952 (19)C5—C4i1.377 (2)
C2—C31.374 (2)N3—N3i1.345 (3)
N1—H1A0.8800N3—H3A0.8800
N1—N21.3457 (19)N3—C41.334 (2)
C3—N21.334 (2)C4—H40.98 (2)
C3—H30.951 (19)
C2—C1—H1130.6 (11)N1—N2—H2125.7
N1—C1—C2108.04 (14)C3—N2—N1108.50 (14)
N1—C1—H1121.4 (11)C3—N2—H2125.7
C1—C2—Cl1127.15 (13)C4i—C5—Cl2127.11 (10)
C3—C2—Cl1126.74 (12)C4—C5—Cl2127.11 (10)
C3—C2—C1106.09 (14)C4—C5—C4i105.8 (2)
C1—N1—H1A125.6N3i—N3—H3A125.7
C1—N1—N2108.87 (13)C4—N3—N3i108.61 (9)
N2—N1—H1A125.6C4—N3—H3A125.7
C2—C3—H3131.0 (12)C5—C4—H4129.2 (11)
N2—C3—C2108.49 (15)N3—C4—C5108.50 (15)
N2—C3—H3120.2 (12)N3—C4—H4122.3 (11)
Cl1—C2—C3—N2178.73 (13)N1—C1—C2—Cl1178.76 (12)
C1—C2—C3—N20.03 (19)N1—C1—C2—C30.01 (18)
C1—N1—N2—C30.07 (18)Cl2—C5—C4—N3179.74 (15)
C2—C1—N1—N20.05 (18)N3i—N3—C4—C50.20 (16)
C2—C3—N2—N10.06 (19)C4i—C5—C4—N30.3 (3)
Symmetry code: (i) x, y+1/2, z.
Hydrogen-bond geometry (Å, º) top
D—H···AD—HH···AD···AD—H···A
N1—H1A···N1ii0.882.032.885 (3)165
N2—H2···N3iii0.881.992.8582 (19)169
N3—H3A···N2iii0.881.992.8582 (19)169
Symmetry codes: (ii) x, y+3/2, z; (iii) x+1, y+1, z.
 

Acknowledgements

The authors would like to acknowledge Dr Indranil Chakraborty and Dr Horst Puschmann for their assistance and crystallographic expertise.

Funding information

Funding for this research was provided by: U.S. Nuclear Regulatory Commission (grant No. NRC-31310018M0012 to Kelly Rue).

References

First citationAlkorta, I., Elguero, J., Donnadieu, B., Etienne, M., Jaffart, J., Schagen, D. & Limbach, H.-H. (1999). New J. Chem. 23, 1231–1237.  Web of Science CSD CrossRef CAS Google Scholar
First citationBruker (2020). APEX3. Bruker AXS Inc., Madison, Wisconsin, USA.  Google Scholar
First citationDolomanov, O. V., Bourhis, L. J., Gildea, R. J., Howard, J. A. K. & Puschmann, H. (2009). J. Appl. Cryst. 42, 339–341.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationFarmiloe, S. E., Berdiell, I. C. & Halcrow, M. E. (2019). CSD Communication (deposition No. 1944671). CCDC, Cambridge, England. https://doi.org/10.5517/ccdc.csd.cc238lbb  Google Scholar
First citationFoces-Foces, C., Llamas-Saiz, A. L. & Elguero, J. (1999). Z. Kristallogr. 214, 237–241.  CAS Google Scholar
First citationGoddard, R., Claramunt, R. M., Escolástico, C. & Elguero, J. (1999). New J. Chem. 23, 237–240.  Web of Science CSD CrossRef CAS Google Scholar
First citationInfantes, L. & Motherwell, S. (2004). Struct. Chem. 15, 173–184.  Web of Science CrossRef CAS Google Scholar
First citationKarrouchi, K., Radi, S., Ramli, Y., Taoufik, J., Mabkhot, Y., Al-aizari, F. A. & Ansar, M. (2018). Molecules, 23, 134.  CrossRef Google Scholar
First citationKrishnakumar, V., Jayamani, N. & Mathammal, R. (2011). Spectrochim. Acta A Mol. Biomol. Spectrosc. 79, 1959–1968.  CrossRef CAS PubMed Google Scholar
First citationKrebs Larsen, F., Lehmann, M. S., Søtofte, I. & Rasmussen, S. E. (1970). Acta Chem. Scand. 24, 3248–3258.  Google Scholar
First citationNaim, M. J., Alam, O., Nawaz, F., Alam, J. & Alam, P. (2016). J. Pharm. Bioallied Sci. 8, 2–17.  CAS PubMed Google Scholar
First citationPérez, J. & Riera, L. (2009). Eur. J. Inorg. Chem. pp. 4913–4925.  Google Scholar
First citationPettinari, C., Masciocchi, N., Pandolfo, L. & Pucci, D. (2010). Chem. Eur. J. 16, 1106–1123.  CrossRef CAS PubMed Google Scholar
First citationRaptis, R. G., Staples, R. J., King, C. & Fackler, J. P. (1993). Acta Cryst. C49, 1716–1719.  CSD CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationRigaku OD (2019). CrysAlis PRO. Rigaku Oxford Diffraction, Yarnton, England.  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 citationViciano-Chumillas, M., Tanase, S., de Jongh, L. J. & Reedijk, J. (2010). Eur. J. Inorg. Chem. pp. 3403–3418.  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