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

Journal logoCRYSTALLOGRAPHIC
COMMUNICATIONS
ISSN: 2056-9890

Synthesis, crystal structure and Hirshfeld surface analysis of the ortho­rhom­bic polymorph of 4-bromo-N-(4-bromo­benzyl­­idene)aniline

crossmark logo

aPG and Research Department of Physics, Arts and Science College, Tiruchirappalli - 620 005, India, bCrystal Growth and Thin Film Laboratory, Department of Physics, Bharathidasan University, Tiruchirappalli - 620 024, India, cUltrafast and Nonlinear Optics Laboratory, Raman Research Institute, C. V. Raman Avenue, Sadashivanagar, Bangalore 560 080, India, and dInstitute of Physics, University of Neuchâtel, rue Emile-Argand 11, 2000 Neuchâtel, Switzerland
*Correspondence e-mail: viji.suba@gmail.com, helen.stoeckli-evans@unine.ch

Edited by W. T. A. Harrison, University of Aberdeen, United Kingdom (Received 9 January 2023; accepted 5 February 2023; online 9 February 2023)

The crystal structure of the title compound, C13H9Br2N [systematic name: (E)-N,1-bis­(4-bromo­phen­yl)methanimine], is a second polymorph (Form II) crystallizing in the ortho­rhom­bic space group Pccn. The first polymorph (Form I) crystallizes in the monoclinic space group P21/c [Bernstein & Izak (1975[Bernstein, J. & Izak, I. (1975). J. Cryst. Mol. Struct. 5, 257-266.]). J. Cryst. Mol. Struct. 5, 257–266; Marin et al. (2013[Marin, L., Harabagiu, V., van der Lee, A., Arvinte, A. & Barboiu, M. (2013). J. Mol. Struct. 1049, 377-385.]). J. Mol. Struct. 1049, 377–385]. The mol­ecule is disordered about an inversion center situated in the middle of the C=N bond, similar to the situation in the monoclinic polymorph: the C=N bond length is 1.243 (7) Å. In the crystal, mol­ecules stack along the b-axis direction and are linked by C—H⋯π inter­actions. The inter­atomic contacts in the crystal for both polymorphs were studied by Hirshfeld surface analysis and have notable differences. The solid-state fluorescence spectrum of Form II shows an emission peak at ca 469 nm. The two-photon absorption coefficient measured from the open aperture Z-scan technique is 1.3 × 10 −11 m W−1, hence, Form II shows optical limiting behaviour.

1. Chemical context

Many compounds possess the ability to form polymorphs (Rolf, 2006[Rolf, H. (2006). Polymorphism in the Pharmaceutical Industry. Wiley: Weinheim, Germany.]; Caira, 2017[Caira, M. R. (2017). Chemistry, Molecular Sciences and Chemical Engineering. Comprehensive Supramolecular Chemistry II, pp. 127-160. Amsterdam: Elsevier.]) and polymorphism and disorder are well established in N-benzyl­ideneaniline compounds. Homo-disubstituted benzyl­ideneaniline compounds such as N-(4-chloro­benzyl­idene)-4-chloro­aniline (Bernstein & Schmidt, 1972[Bernstein, J. & Schmidt, G. M. J. (1972). J. Chem. Soc. Perkin Trans. 2, pp. 951-955.]; Bernstein & Izak, 1976[Bernstein, J. & Izak, I. (1976). J. Chem. Soc. Perkin Trans. 2, pp. 429-434.]) and N-(4-methyl­benzyl­idene)-4-methyl­aniline (Bar & Bernstein, 1977[Bar, I. & Bernstein, J. (1977). Acta Cryst. B33, 1738-1744.], 1982[Bar, I. & Bernstein, J. (1982). Acta Cryst. B38, 121-125.]; Bernstein & Izak,1976[Bernstein, J. & Izak, I. (1976). J. Chem. Soc. Perkin Trans. 2, pp. 429-434.]) exist in dimorphic and trimorphic forms, respectively. Bar & Bernstein (1983[Bar, I. & Bernstein, J. (1983). Acta Cryst. B39, 266-272.]) reported a detailed description of three types of disorder in N-benzyl­ideneanilines, viz., positional, orientational and substitutional. These various types of disorder have been discussed in relation to hetero-disubstituted 4-X-N-(4′-nitro­benzl­idene)anilines (where X = H, F, Cl, Br, CH3, CH3O, OH) by Leela et al. (2020[Leela, S., Subashini, A., Reji, P., Ramamurthi, K. & Stoeckli-Evans, H. (2020). Acta Cryst. E76, 417-422.] and references therein).

[Scheme 1]

The crystal structure of the monoclinic polymorph of the title compound, N-4-bromo­benzyl­idene)-4-bromo­aniline, C13H9Br2N (Form I) has been reported by Bernstein & Izak (1975[Bernstein, J. & Izak, I. (1975). J. Cryst. Mol. Struct. 5, 257-266.]) and later by Marin et al. (2013[Marin, L., Harabagiu, V., van der Lee, A., Arvinte, A. & Barboiu, M. (2013). J. Mol. Struct. 1049, 377-385.]). Bernstein & Hagler (1979[Bernstein, J. & Hagler, A. T. (1979). Mol. Cryst. Liq. Cryst. 50, 223-233.]) stated that they did not observe polymorphism for this compound. However, Marin et al. (2013[Marin, L., Harabagiu, V., van der Lee, A., Arvinte, A. & Barboiu, M. (2013). J. Mol. Struct. 1049, 377-385.]) reported that a second polymorph could appear during thermal treatment.

In the present study the homo-disubstituted title compound was synthesized using the procedure described by Bernstein & Izak (1975[Bernstein, J. & Izak, I. (1975). J. Cryst. Mol. Struct. 5, 257-266.]). Single crystals were grown by slow evaporation of a solution in either ethanol or a mixture of methanol/chloro­form (1:1). Crystals from both essays proved to be those of the ortho­rhom­bic polymorph – Form II.

2. Structural commentary

The title compound crystallizes in the ortho­rhom­bic space group Pccn (Form II). It exhibits disorder about a crystallographic inversion center situated on the middle of the C7=N1 bond, as illustrated in Fig. 1[link]. This arrangement is similar to that observed for the monoclinic polymorph (Bernstein & Izak, 1975[Bernstein, J. & Izak, I. (1975). J. Cryst. Mol. Struct. 5, 257-266.]; Marin et al., 2013[Marin, L., Harabagiu, V., van der Lee, A., Arvinte, A. & Barboiu, M. (2013). J. Mol. Struct. 1049, 377-385.]) and for the ortho­rhom­bic and monoclinic polymorphs of the chloro-disubstituted N-benzyl­ideneaniline (Bernstein & Izak, 1976[Bernstein, J. & Izak, I. (1976). J. Chem. Soc. Perkin Trans. 2, pp. 429-434.]; Bernstein & Schmidt, 1972[Bernstein, J. & Schmidt, G. M. J. (1972). J. Chem. Soc. Perkin Trans. 2, pp. 951-955.]).

[Figure 1]
Figure 1
A view of the mol­ecular structure of Form II of the title compound, with atom labelling. Displacement ellipsoids are drawn at the 50% probability level. Symmetry code: (i) −x, −y, −z.

Atoms C7, H7 and N1 were refined with occupancies of 0.5 each. The C=N bond length is 1.243 (7) Å, and the benzene rings are co-planar by symmetry. These geometrical parameters are similar to those observed for Form I, the monoclinic polymorph (Marin et al., 2013[Marin, L., Harabagiu, V., van der Lee, A., Arvinte, A. & Barboiu, M. (2013). J. Mol. Struct. 1049, 377-385.]).

A search of the Cambridge Structural Database [CSD, V5.43, last update November 2022; Groom et al., (2016[Groom, C. R., Bruno, I. J., Lightfoot, M. P. & Ward, S. C. (2016). Acta Cryst. B72, 171-179.])] for N-benzyl­ideneanilines with no ortho-substituents on the aromatic rings, no errors, no polymerics, no ions or disorder, for organics only and R < 0.05 gave 220 hits. An analysis of the geometry of these 220 compounds (Mercury; Macrae et al., 2020[Macrae, C. F., Sovago, I., Cottrell, S. J., Galek, P. T. A., McCabe, P., Pidcock, E., Platings, M., Shields, G. P., Stevens, J. S., Towler, M. & Wood, P. A. (2020). J. Appl. Cryst. 53, 226-235.]) indicated that the C=N bond length varies from 1.1775 to 2.202 Å, with a mean value of 1.269 Å (mean deviation of 0.009 Å). Hence, the value observed for Form II is significantly shorter at 1.243 (7) Å, while that for Form I is close to this average at 1.257 (2) Å. The bond lengths N1—C1ar and C7—C1ar have mean values of 1.420 (mean deviation of 0.006 Å) and 1.462 Å (mean deviation of 0.008 Å), respectively. The values observed for Form I are the same at 1.439 (3) Å, and for Form II they are also the same at 1.45 (2) Å, owing to the fact that both polymorphs are disordered about an inversion centre.

3. Supra­molecular features

In the crystal, the mol­ecules are linked by C—H⋯π inter­actions (Table 1[link]), forming columns stacking along the b-axis direction, as shown in Fig. 2[link].

Table 1
Hydrogen-bond geometry (Å, °)

Cg1 is the centroid of the C1–C6 ring.

D—H⋯A D—H H⋯A DA D—H⋯A
C2—H2⋯Cg1i 0.95 2.81 3.535 (3) 134
C5—H5⋯Cg1ii 0.95 2.78 3.494 (3) 133
Symmetry codes: (i) [x, -y-{\script{1\over 2}}, z-{\script{1\over 2}}]; (ii) [x, -y+{\script{1\over 2}}, z+{\script{1\over 2}}].
[Figure 2]
Figure 2
A view along the c-axis of the crystal packing of Form II of the title compound.

4. Hirshfeld surface analyses and two-dimensional fingerprint plots for Form I and Form II

The Hirshfeld surface analysis (Spackman & Jayatilaka, 2009[Spackman, M. A. & Jayatilaka, D. (2009). CrystEngComm, 11, 19-32.]) was performed and the associated two-dimensional fingerprint plots (McKinnon et al., 2007[McKinnon, J. J., Jayatilaka, D. & Spackman, M. A. (2007). Chem. Commun. pp. 3814-3816.]) were generated with CrystalExplorer17 (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.]) following the protocol of Tan et al. (2019[Tan, S. L., Jotani, M. M. & Tiekink, E. R. T. (2019). Acta Cryst. E75, 308-318.]). The Hirshfeld surface is colour-mapped with the normalized contact distance, dnorm, from red (distances shorter than the sum of the van der Waals radii) through white to blue (distances longer than the sum of the van der Waals radii).

The Hirshfeld surfaces of Form I and Form II mapped over dnorm, are given in Fig. 3[link]. The faint red spots indicate that short contacts are significant in the crystal packing of both compounds.

[Figure 3]
Figure 3
(a) The Hirshfeld surface of Form I, mapped over dnorm in the colour range −0.0139 to 0.7999 a.u. and (b) the Hirshfeld surface of Form II, mapped over dnorm in the colour range −0.0329 to 1.0662 a.u.

The full two-dimensional fingerprint plots for Form I and Form II are given in Figs. 4[link] and 5[link], respectively. There it can be seen that the relative contributions of the various inter­atomic contacts in the two polymorphs are significantly different (Table 2[link]). For example, the C⋯H/H⋯C contacts have a major contribution (36.4%) in Form II whereas the same contact type is only 15.3% in Form I. This trend is reversed for the Br⋯H/H⋯Br contacts, which make a major contribution (33.0%) in Form I but less, 24.9%, in Form II, and for the H⋯H contacts which are greater in Form I (30.8%) compared to Form II (21.2%). The halogen⋯halogen inter­actions at the para-positions of the aromatic rings are more important in Form II (9.7%) compared to Form I (4.1%). Similar values of relative percentage are observed for the Br⋯C/C⋯Br contacts in Form I (2.6%) and Form II (2.3%). There are significant C⋯C inter­molecular contacts observed for Form I (9.4%) but they are very weak (0.1%) in Form II. In contrast, the N⋯H/H⋯N contacts observed for Form II (5.4) are stronger than those observed for Form I (3.7%).

Table 2
The relative contributions of the various inter­atomic contacts (Å) in the crystal structures of Form Ia and Form II

Contact Form Ia Form II
  % %
H⋯H 30.8 21.2
C⋯H/H⋯C 15.3 36.4
N⋯H/H⋯N 3.7 5.4
Br⋯H/H⋯Br 33.0 24.9
C⋯C 9.4 0.1
Br⋯C/C⋯Br 2.6 2.3
Br⋯Br 4.1 9.7
Note: (a) Marin et al. (2013[Marin, L., Harabagiu, V., van der Lee, A., Arvinte, A. & Barboiu, M. (2013). J. Mol. Struct. 1049, 377-385.]).
[Figure 4]
Figure 4
(a) The full two-dimensional fingerprint plot for Form I, and those delineated into Br⋯H/H⋯Br, H⋯H, C⋯H/H⋯C, C⋯C, Br⋯Br, N⋯H/H⋯N and Br⋯C/C⋯Br and contacts.
[Figure 5]
Figure 5
(a) The full two-dimensional fingerprint plot for Form II, and those delineated into C⋯H/H⋯C, Br⋯H/H⋯Br, H⋯H, Br⋯Br, N⋯H/H⋯N, Br⋯C/C⋯Br and C⋯C contacts.

5. Thermal properties

Thermogravimetric (TGA) and differential thermal (DTA) analyses were recorded using an S·T.A. − 1500 Simultaneous Thermo Analytical system in the temperature region 30–300°C at a heating rate of 10 K min−1 under a nitro­gen atmosphere and alumina (Al2O3) was used as the reference material. The thermogravimetric and differential thermal analyses of Form II are shown in Fig. 6[link]. In the DTA curve the sharp endothermic peak observed at ca. 145 °C corresponds to the melting point and indicates that there is no decomposition before melting. A single-stage weight loss is indicated in the TGA curve. At ca 300°C almost the entire mass of the sample is lost, thus indicating the occurrence of bulk decomposition.

[Figure 6]
Figure 6
The TGA (blue) and DTA (black) curves for Form II.

6. Fluorescence spectra

The fluorescence emission spectrum of Form II at room temperature was recorded using a Horiba Jobin Yuon FLUOROLOG-FL3-11 spectro­fluoro­meter in the wavelength range of 375–600 nm (Fig. 7[link]). A powdered sample of Form II was excited at 375 nm and an emission peak was observed at 469 nm (blue colour) due to the presence of the aromatic rings (Lakowicz, 2006[Lakowicz, J. R. (2006). Principles of Fluorescence Spectroscopy, 3rd ed. New York: Springer.]). For Form I, an emission peak was reported at 414 nm (weak violet colour) when excited at 278 nm (Marin et al., 2013[Marin, L., Harabagiu, V., van der Lee, A., Arvinte, A. & Barboiu, M. (2013). J. Mol. Struct. 1049, 377-385.]).

[Figure 7]
Figure 7
The solid-state fluorescence emission spectrum of Form II.

7. Z-scan studies

Employing the open aperture Z-scan technique the third order non-linear optical property of Form II was studied: the experimental setup has been described previously (Subashini et al., 2013a[Subashini, A., Bhagavannarayana, G. & Ramamurthi, K. (2013a). Spectrochim. Acta A Mol. Biomol. Spectrosc. 104, 403-408.],b[Subashini, A., Leela, S., Ramamurthi, K., Arakcheeva, A., Stoeckli-Evans, H., Petříček, V., Chapuis, G., Pattison, P. & Reji, P. (2013b). CrystEngComm, 15, 2474-2481.]). A sample of Form II was dissolved in chloro­form, and 5 ns laser pulses at 532 nm were used for excitation. The optical density of the Form II solution is low at 532 nm, and its linear transmission is high (85%). The laser pulse energy reaching the sample was 100 µJ. The open-aperture transmission normalized to the linear transmission of the sample (normalized transmittance) was then plotted against the sample position measured relative to the beam focus and the non-linear absorption was indicated by a smooth valley-shaped curve, which was symmetric about the focal (Z = 0) position. The Z scan of pure chloro­form was run separately to ensure that the solvent showed no non-linearity under the same experimental conditions.

The results of the open-aperture Z scan of Form II are shown in Fig. 8[link]. An increase in absorption is found as the laser intensity is increased, thus indicating the non-linear optical absorption and the optical limiting behaviour of Form II. Since the linear transmission of the sample is ∼85%, the observed non-linear absorption might have contributions from genuine two-photon absorption (2PA), as well as from excited state absorption (ESA). For further details of the data analysis, see supporting information.

[Figure 8]
Figure 8
The open-aperture Z scan of Form II.

8. Synthesis, crystallization and spectroscopic analyses

The title compound was synthesized by the condensation reaction of 4-bromo­benzaldehyde with 4-bromo­aniline in an equimolar ratio: the reactants were dissolved in ethanol and refluxed for 6 h at 363 K, then cooled to room temperature. The precipitated product was purified by repeated recrystallization using ethanol as solvent. Colourless crystals of the title compound were grown at room temperature by slow evaporation of a solution in either ethanol or a mixture of methanol/chloro­form (1:1).

The solid-state Fourier Transform Infrared (FT–IR) and Fourier Transform Raman (FT–Raman) spectra at room temperature (Fig. 9[link]) were recorded using a Perkin Elmer grating infrared spectrophotometer (KBr pellet technique) and a Varian FT-Raman spectrometer, respectively, in the wave number range of 400–4000 cm−1 (Fig. 11). The C=N stretching vibration is observed as a strong band at 1630 cm−1 (IR) and at 1615 cm−1 (Raman) (Silverstein et al., 2005[Silverstein, M., Webster, X. & Kiemle, J. (2005). Spectrometric Identification of Organic Compounds. Wiley, New York.]). For Form I, the C=N stretching vibration was reported at 1620 cm−1 (Marin et al., 2013[Marin, L., Harabagiu, V., van der Lee, A., Arvinte, A. & Barboiu, M. (2013). J. Mol. Struct. 1049, 377-385.]). In the IR spectrum, the aromatic C=C stretching vibrations are observed at 1468 and 1562 cm−1, and at 1474 and 1558 cm−1 in the Raman spectrum. In the IR spectrum, the aromatic C—H in-plane bending modes are observed at 1059, 1111, 1163 and 1230 cm−1 whereas the out-of-plane bending modes appear at 819 and 1000 cm−1 (Yoshino et al., 2013[Yoshino, J., Kano, N. & Kawashima, T. (2013). Dalton Trans. 42, 15826-15834.]). In the Raman spectrum, the corresponding in-plane bending modes are observed at 1008, 1067, 1098, 1165 and 1183 cm−1 whereas the out-of-plane bending mode is observed at 879 cm−1. In the FT–IR spectrum, the band at 521 cm−1 is due to the aromatic C—Br stretching vibration.

[Figure 9]
Figure 9
The FT–IR (black) and FT–Raman (blue) spectra of Form II.

The 1H and 13C NMR data for Form II in chloro­form-D (CDCl3) using tetra­methyl­silane as the inter­nal standard were recorded employing a Bruker AC 400-NMR spectrometer.

The 1H NMR spectrum (illustrated in Fig. S1 of the supporting information) exhibits five proton signals. The intense proton signal appearing at δ = 8.364 ppm is attributed to the imine group. The peaks at 7.48–7.52 and 7.59–7.61 ppm are due to the aromatic ring proton signals of the bromo­benzaldehyde moiety and the peaks at 7.07–7.09 and 7.74–7.76 ppm are due to the aromatic ring proton signals of the bromo­aniline moiety.

In the 13C NMR spectrum (illustrated in Fig. S2 of the supporting information) the peak at 76.46–77.40 ppm corres­ponds to the carbon atom of CDCl3. A peak corresponding to the imine carbon atom is observed at δ = 159.26 ppm. The peaks at 119.60, 122.54, 134.88 and 150.54 ppm correspond to the phenyl carbon-atom signals of the bromo­aniline moiety and the peaks at 126.19, 126.24, 130.19, 130.24, 132.09, 132.13, 132.24 and 132.29 ppm are due to phenyl carbon-atom signals of the bromo­benzaldehyde moiety.

9. Refinement details

Crystal data, data collection and structure refinement details are summarized in Table 3[link]. The atoms N1, C7, and H7 were refined with occupancies of 0.5 each. The C-bound H atoms were included in calculated positions (C—H = 0.95 Å) and refined as riding atoms with Uiso(H) = 1.2Ueq(C).

Table 3
Experimental details

Crystal data
Chemical formula C13H9Br2N
Mr 339.03
Crystal system, space group Orthorhombic, Pccn
Temperature (K) 173
a, b, c (Å) 27.4096 (19), 7.3301 (5), 5.9135 (3)
V3) 1188.11 (13)
Z 4
Radiation type Mo Kα
μ (mm−1) 6.79
Crystal size (mm) 0.45 × 0.33 × 0.13
 
Data collection
Diffractometer STOE IPDS 2
Absorption correction Multi-scan (MULABS; Spek, 2020[Spek, A. L. (2020). Acta Cryst. E76, 1-11.])
Tmin, Tmax 0.411, 1.000
No. of measured, independent and observed [I > 2σ(I)] reflections 11334, 1125, 911
Rint 0.075
(sin θ/λ)max−1) 0.609
 
Refinement
R[F2 > 2σ(F2)], wR(F2), S 0.037, 0.061, 1.09
No. of reflections 1125
No. of parameters 71
H-atom treatment H-atom parameters constrained
Δρmax, Δρmin (e Å−3) 0.34, −0.48
Computer programs: X-AREA and X-RED32 (Stoe & Cie, 2009[Stoe & Cie (2009). X-AREA and X-RED32. Stoe & Cie GmbH, Darmstadt, Germany.]), SHELXS97 (Sheldrick, 2008[Sheldrick, G. M. (2008). Acta Cryst. A64, 112-122.]), SHELXL2018/3(Sheldrick, 2015[Sheldrick, G. M. (2015). Acta Cryst. C71, 3-8.]), PLATON (Spek, 2020[Spek, A. L. (2020). Acta Cryst. E76, 1-11.]), Mercury (Macrae et al., 2020[Macrae, C. F., Sovago, I., Cottrell, S. J., Galek, P. T. A., McCabe, P., Pidcock, E., Platings, M., Shields, G. P., Stevens, J. S., Towler, M. & Wood, P. A. (2020). J. Appl. Cryst. 53, 226-235.]) and publCIF (Westrip, 2010[Westrip, S. P. (2010). J. Appl. Cryst. 43, 920-925.]).

Supporting information


Computing details top

Data collection: X-AREA (Stoe & Cie, 2009); cell refinement: X-AREA (Stoe & Cie, 2009); data reduction: X-RED32 (Stoe & Cie, 2009); program(s) used to solve structure: SHELXS97 (Sheldrick, 2008); program(s) used to refine structure: SHELXL2018/3(Sheldrick, 2015); molecular graphics: PLATON (Spek, 2020) and Mercury (Macrae et al., 2020); software used to prepare material for publication: SHELXL2018/3 (Sheldrick, 2015), PLATON (Spek, 2020) and publCIF (Westrip, 2010).

(E)-N,1-Bis(4-bromophenyl)methanimine top
Crystal data top
C13H9Br2NDx = 1.895 Mg m3
Mr = 339.03Mo Kα radiation, λ = 0.71073 Å
Orthorhombic, PccnCell parameters from 8006 reflections
a = 27.4096 (19) Åθ = 1.5–26.1°
b = 7.3301 (5) ŵ = 6.79 mm1
c = 5.9135 (3) ÅT = 173 K
V = 1188.11 (13) Å3Plate, colourless
Z = 40.45 × 0.33 × 0.13 mm
F(000) = 656
Data collection top
STOE IPDS 2
diffractometer
1125 independent reflections
Radiation source: fine-focus sealed tube911 reflections with I > 2σ(I)
Plane graphite monochromatorRint = 0.075
φ + ω scansθmax = 25.7°, θmin = 1.5°
Absorption correction: multi-scan
(MULABS; Spek, 2020)
h = 3333
Tmin = 0.411, Tmax = 1.000k = 88
11334 measured reflectionsl = 67
Refinement top
Refinement on F2Secondary atom site location: difference Fourier map
Least-squares matrix: fullHydrogen site location: inferred from neighbouring sites
R[F2 > 2σ(F2)] = 0.037H-atom parameters constrained
wR(F2) = 0.061 w = 1/[σ2(Fo2) + (0.0216P)2 + 1.2266P]
where P = (Fo2 + 2Fc2)/3
S = 1.09(Δ/σ)max < 0.001
1125 reflectionsΔρmax = 0.34 e Å3
71 parametersΔρmin = 0.48 e Å3
0 restraintsExtinction correction: (SHELXL-2018/3; Sheldrick, 2015), Fc*=kFc[1+0.001xFc2λ3/sin(2θ)]-1/4
Primary atom site location: structure-invariant direct methodsExtinction coefficient: 0.0018 (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*/UeqOcc. (<1)
Br10.20539 (2)0.04724 (6)0.54116 (7)0.03603 (15)
C10.06221 (13)0.0182 (5)0.1208 (6)0.0249 (7)
C20.10666 (12)0.0797 (4)0.0348 (6)0.0263 (8)
H20.1075470.1368720.1093460.032*
C30.14960 (12)0.0586 (5)0.1562 (6)0.0269 (8)
H30.1798250.0986880.0952360.032*
C40.14765 (12)0.0211 (5)0.3664 (6)0.0234 (7)
C50.10402 (12)0.0822 (4)0.4573 (6)0.0241 (7)
H50.1033750.1372530.6026600.029*
C60.06141 (12)0.0622 (5)0.3344 (6)0.0247 (8)
H60.0313730.1035670.3960990.030*
N10.0189 (6)0.0524 (16)0.013 (3)0.0249 (7)0.5
C70.0200 (8)0.0280 (19)0.026 (4)0.0249 (7)0.5
H70.0239690.0809040.1719460.037*0.5
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
Br10.02284 (19)0.0454 (2)0.0398 (2)0.00407 (19)0.00786 (17)0.0014 (2)
C10.0279 (12)0.0196 (18)0.0272 (15)0.0043 (13)0.0033 (11)0.0045 (13)
C20.0331 (18)0.0231 (19)0.0229 (17)0.0017 (14)0.0001 (15)0.0011 (16)
C30.0257 (16)0.028 (2)0.0267 (19)0.0032 (17)0.0051 (14)0.0001 (17)
C40.0196 (15)0.0229 (18)0.0276 (18)0.0043 (14)0.0001 (13)0.0045 (15)
C50.0287 (17)0.0206 (19)0.0229 (16)0.0023 (13)0.0026 (15)0.0005 (15)
C60.0191 (15)0.0226 (19)0.032 (2)0.0011 (15)0.0016 (14)0.0002 (17)
N10.0279 (12)0.0196 (18)0.0272 (15)0.0043 (13)0.0033 (11)0.0045 (13)
C70.0279 (12)0.0196 (18)0.0272 (15)0.0043 (13)0.0033 (11)0.0045 (13)
Geometric parameters (Å, º) top
Br1—C41.900 (3)C3—H30.9500
C1—C61.394 (5)C4—C51.385 (5)
C1—C21.395 (5)C5—C61.383 (5)
C1—N11.448 (17)C5—H50.9500
C1—C71.45 (2)C6—H60.9500
C2—C31.387 (5)N1—N1i1.30 (3)
C2—H20.9500C7—C7i1.21 (4)
C3—C41.374 (5)C7—H70.9500
C6—C1—C2118.7 (3)C3—C4—Br1120.1 (3)
C6—C1—N1123.8 (8)C5—C4—Br1118.4 (3)
C2—C1—N1117.4 (7)C6—C5—C4119.4 (3)
C6—C1—C7123.6 (9)C6—C5—H5120.3
C2—C1—C7117.5 (9)C4—C5—H5120.3
C3—C2—C1121.1 (3)C5—C6—C1120.5 (3)
C3—C2—H2119.5C5—C6—H6119.8
C1—C2—H2119.5C1—C6—H6119.8
C4—C3—C2118.8 (3)N1i—N1—C1119.0 (18)
C4—C3—H3120.6C7i—C7—C1123 (3)
C2—C3—H3120.6C7i—C7—H7118.3
C3—C4—C5121.5 (3)C1—C7—H7118.3
C6—C1—C2—C31.3 (5)C4—C5—C6—C10.1 (5)
N1—C1—C2—C3177.6 (5)C2—C1—C6—C50.7 (5)
C7—C1—C2—C3173.6 (6)N1—C1—C6—C5176.7 (6)
C1—C2—C3—C41.3 (5)C7—C1—C6—C5173.9 (6)
C2—C3—C4—C50.7 (5)C6—C1—N1—N1i28.0 (17)
C2—C3—C4—Br1178.5 (3)C2—C1—N1—N1i155.9 (13)
C3—C4—C5—C60.1 (5)C6—C1—C7—C7i2 (2)
Br1—C4—C5—C6179.1 (3)C2—C1—C7—C7i177.0 (15)
Symmetry code: (i) x, y, z.
Hydrogen-bond geometry (Å, º) top
Cg1 is the centroid of the C1–C6 ring.
D—H···AD—HH···AD···AD—H···A
C2—H2···Cg1ii0.952.813.535 (3)134
C5—H5···Cg1iii0.952.783.494 (3)133
Symmetry codes: (ii) x, y1/2, z1/2; (iii) x, y+1/2, z+1/2.
The relative contributions of the various interatomic contacts (Å) in the crystal structures of Form Ia and Form II top
ContactForm IaForm II
%%
H···H30.821.2
C···H/H···C15.336.4
N···H/H···N3.75.4
Br···H/H···Br33.024.9
C···C9.40.1
Br···C/C···Br2.62.3
Br···Br4.19.7
Note: (a) Marin et al. (2013).
 

Acknowledgements

The authors thank the Central Instrumentation Facility, Pondicherry University, and the Advanced Instrumentation Research Facility, Jawaharlal Nehru University, New Delhi, for access to their analytical facilities. HSE is grateful to the University of Neuchâtel for their support over the years.

Funding information

AS thanks the University Grants Commission, New Delhi, for the award of a UGC Meritorious Fellowship [File No. 4-1/2008 (BRS)].

References

First citationBar, I. & Bernstein, J. (1977). Acta Cryst. B33, 1738–1744.  CSD CrossRef CAS IUCr Journals Web of Science Google Scholar
First citationBar, I. & Bernstein, J. (1982). Acta Cryst. B38, 121–125.  CSD CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationBar, I. & Bernstein, J. (1983). Acta Cryst. B39, 266–272.  CSD CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationBernstein, J. & Hagler, A. T. (1979). Mol. Cryst. Liq. Cryst. 50, 223–233.  CrossRef CAS Web of Science Google Scholar
First citationBernstein, J. & Izak, I. (1975). J. Cryst. Mol. Struct. 5, 257–266.  CSD CrossRef CAS Web of Science Google Scholar
First citationBernstein, J. & Izak, I. (1976). J. Chem. Soc. Perkin Trans. 2, pp. 429–434.  CSD CrossRef Web of Science Google Scholar
First citationBernstein, J. & Schmidt, G. M. J. (1972). J. Chem. Soc. Perkin Trans. 2, pp. 951–955.  CSD CrossRef Web of Science Google Scholar
First citationCaira, M. R. (2017). Chemistry, Molecular Sciences and Chemical Engineering. Comprehensive Supramolecular Chemistry II, pp. 127–160. Amsterdam: Elsevier.  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 citationLakowicz, J. R. (2006). Principles of Fluorescence Spectroscopy, 3rd ed. New York: Springer.  Google Scholar
First citationLeela, S., Subashini, A., Reji, P., Ramamurthi, K. & Stoeckli-Evans, H. (2020). Acta Cryst. E76, 417–422.  Web of Science CSD CrossRef IUCr Journals Google Scholar
First citationMacrae, C. F., Sovago, I., Cottrell, S. J., Galek, P. T. A., McCabe, P., Pidcock, E., Platings, M., Shields, G. P., Stevens, J. S., Towler, M. & Wood, P. A. (2020). J. Appl. Cryst. 53, 226–235.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationMarin, L., Harabagiu, V., van der Lee, A., Arvinte, A. & Barboiu, M. (2013). J. Mol. Struct. 1049, 377–385.  Web of Science CSD CrossRef CAS Google Scholar
First citationMcKinnon, J. J., Jayatilaka, D. & Spackman, M. A. (2007). Chem. Commun. pp. 3814–3816.  Web of Science CrossRef Google Scholar
First citationRolf, H. (2006). Polymorphism in the Pharmaceutical Industry. Wiley: Weinheim, Germany.  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 citationSilverstein, M., Webster, X. & Kiemle, J. (2005). Spectrometric Identification of Organic Compounds. Wiley, New York.  Google Scholar
First citationSpackman, M. A. & Jayatilaka, D. (2009). CrystEngComm, 11, 19–32.  Web of Science CrossRef CAS 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 citationSpek, A. L. (2020). Acta Cryst. E76, 1–11.  Web of Science CrossRef IUCr Journals Google Scholar
First citationStoe & Cie (2009). X-AREA and X-RED32. Stoe & Cie GmbH, Darmstadt, Germany.  Google Scholar
First citationSubashini, A., Bhagavannarayana, G. & Ramamurthi, K. (2013a). Spectrochim. Acta A Mol. Biomol. Spectrosc. 104, 403–408.  Web of Science CrossRef CAS PubMed Google Scholar
First citationSubashini, A., Leela, S., Ramamurthi, K., Arakcheeva, A., Stoeckli-Evans, H., Petříček, V., Chapuis, G., Pattison, P. & Reji, P. (2013b). CrystEngComm, 15, 2474–2481.  Web of Science CSD CrossRef CAS Google Scholar
First citationTan, S. L., Jotani, M. M. & Tiekink, E. R. T. (2019). Acta Cryst. E75, 308–318.  Web of Science CrossRef IUCr Journals Google Scholar
First citationWestrip, S. P. (2010). J. Appl. Cryst. 43, 920–925.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationYoshino, J., Kano, N. & Kawashima, T. (2013). Dalton Trans. 42, 15826–15834.  Web of Science CrossRef CAS PubMed 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