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

Journal logoCRYSTALLOGRAPHIC
COMMUNICATIONS
ISSN: 2056-9890

Synthesis, crystal structure, Hirshfeld surface analysis and DFT study of the 1,1′-(buta-1,3-diyne-1,4-di­yl)bis­­(cyclo­hexan-1-ol)

crossmark logo

aChirchik State Pedagogical University, 111700, A. Temur Str. 104, Chirchik, Uzbekistan, and bInstitute of Bioorganic Chemistry, Academy of Sciences of Uzbekistan, 100125, M. Ulugbek Str 83, Tashkent, Uzbekistan
*Correspondence e-mail: atom.uz@mail.ru

Edited by A. S. Batsanov, University of Durham, United Kingdom (Received 28 November 2022; accepted 31 May 2023; online 2 June 2023)

The title compound, C16H22O2, was synthesized in order to obtain its guest-free form because `wheel-and-axle'-shaped mol­ecules tend to crystallize from solutions as solvates or host–guest mol­ecules. It crystallizes in the monoclinic space group P2/c with two crystallographically non-equivalent mol­ecules, one situated on an inversion center and the other on a twofold axis. The rod-like 1,3-diyne fragments have the usual linear geometry. In the crystal, O—H ⋯ O bonds form eight-membered rings of the R44(8) type, linking mol­ecules into layers. The Hirshfeld surface analysis indicates that the largest con­tributions are from inter­molecular H⋯H (ca 71%) and H⋯C/C⋯H (ca 19%) contacts. The energies of the frontier mol­ecular orbitals were determined by DFT calculations at the B3LYP/def2-TZVP level of theory.

1. Chemical context

The presence of two triple C≡C bonds and two hy­droxy groups in the mol­ecules of di­acetyl­ene diols R1R2(OH)C—C≡C—C≡C—C(OH)R3R4, as well as substituents with different structures and functional groups containing heteroatoms, increases the possibilities of synthesis and the production of valuable, chemically stable and biologically active compounds based on such compounds (Cadierno, 2022[Cadierno, V. (2022). Catalysts, 12, 2-48.]). In particular, as the hydrogen atom adjacent to the strong C≡C bond is labile (Brücner, 2010[Brücner, R. (2010). Organic Mechanisms: Reactions, Stereochemistry and Synthesis, 3rd ed, pp. 6-8. Berlin, Heidelberg: Springer-Verlag.]), terminal alkynes easily undergo nucleophilic addition reactions to the carbonyl group and terminal (Hosseini et al., 2020[Hosseini, A. & Schreiner, P. R. (2020). Eur. J. Org. Chem. pp. 4339-4346.]; Sum et al., 2013[Sum, Y. N., Yu, D. & Zhang, Y. (2013). Green Chem. 15, 2718-2721.]) or inter­nal acetyl­ene alcohols (Tanaka et al., 2011[Tanaka, K., Kukita, K., Ichibakase, T., Kotani, S. & Nakajima, M. (2011). Chem. Commun. 47, 5614-5616.]; Motoki et al., 2007[Motoki, R., Kanai, M. & Shibasaki, M. (2007). Org. Lett. 9, 2997-3000.]) and diols (Ardila-Fierro et al., 2019[Ardila-Fierro, K. J., Bolm, C. & Hernández, J. G. (2019). Angew. Chem. Int. Ed. 58, 12945-12949.]) with various substituents. Di­acetyl­ene diols and polyacetyl­ene diols (Shi Shun et al., 2006[Shi Shun, A. L. K. & Tykwinski, R. R. (2006). Angew. Chem. Int. Ed. 45, 1034-1057.]) can also be synthesized by performing dimerization processes. Many reactions, such as cyclization (Zhang et al., 2010[Zhang, X., Teo, W. T., Sally & Chan, P. W. H. (2010). J. Org. Chem. 75, 6290-6293.]) or substitution (Kuang et al., 2018[Kuang, Z., Chen, H., Yan, J., Yang, K., Lan, Y. & Song, Q. (2018). Org. Lett. 20, 5153-5157.]), based on the hy­droxy group (–OH) or its hydrogen atom in an acetyl­ene alcohol, give opportunities to synthesize new biologically active substances. Hexa-2,4-diene-1,6-diol and its derivatives have been found to have anti­cancer chemotherapeutic properties (Lee et al., 2015[Lee, C.-Y., Yun, J. H., Kang, K., Nho, C.-W. & Shin, D. (2015). Bioorg. Med. Chem. Lett. 25, 4020-4023.]). Moreover, some di­acetyl­ene diols and their derivatives have anti­bacterial (Ankisetty et al., 2012[Ankisetty, S. & Slattery, M. (2012). Mar. Drugs, 10, 1037-1043.]), anti­viral (Geng et al., 2015[Geng, C.-A., Huang, X.-Y., Chen, X.-L., Ma, Y.-B., Rong, G.-Q., Zhao, Y., Zhang, X.-M. & Chen, J.-J. (2015). J. Ethnopharmacol. 176, 109-117.]) and neuritogenic (Wang et al., 2011[Wang, Y.-H., Liu, H., Zhu, L.-L., Li, X.-X. & Chen, Z. (2011). Adv. Synth. Catal. 353, 707-712.]) activities.

Moreover, the above indicated substances behave as versatile host compounds accommodating many guest species (Weber et al., 1991[Weber, E., Atwood, J. L., Davies, J. E. D. & MacNicol, D. D. (1991). Inclusion Compounds, 5th ed, pp. 188-263. Oxford University Press.]) because the shape of their mol­ecules is inefficient for close packing in crystals. Therefore, the preparation of such compounds in their pure form, i.e. a guest-free state, is of inter­est. This paper describes the preparation (Fig. 1[link]), mol­ecular and crystal structure, as well Hirshfeld surface analysis of the guest-free crystal of the title compound, (I)[link].

[Scheme 1]
[Figure 1]
Figure 1
Synthesis of compound (I)[link].

2. Structural commentary

There are the principles of directed host design formulated by Weber (Weber et al., 1991[Weber, E., Atwood, J. L., Davies, J. E. D. & MacNicol, D. D. (1991). Inclusion Compounds, 5th ed, pp. 188-263. Oxford University Press.]), according to which bulky and rigid compounds are packed in crystals inefficiently, leaving suitable cavities for the accommodation of guest mol­ecules. Indeed, host compounds with a `wheel-and-axle' shape of the mol­ecule easily include several guests (Weber et al., 2004[Weber, E., Korkas, P. P., Czugler, M. & Seichter, W. (2004). Supramol. Chem. 16, 217-226.]). However, in the case of compound (I)[link] belonging to this family, only one inclusion compound (with 1,4-di­aza­bicyclo-[2.2.2]octane as the guest) has been structurally characterized (Chandrasekhar et al., 2013[Chandrasekhar, V. & Narayanan, R. S. (2013). Indian J. Chem. Sect. A Inorg. Bio-inorg. Phys. Theor. Anal. Chem. 52, 1066-1072.]). In our experimental conditions we have obtained guest-free crystals of (I)[link]. They belong to the monoclinic system with space group P2/c. There are two crystallographically non-equivalent mol­ecules, both situated on symmetry elements: mol­ecule A is located on an inversion center while mol­ecule B lies on a twofold axis. Thus there are two half-mol­ecules in the asymmetric part of the unit cell. The rod-like 1,3-diyne fragment has the usual linear geometry and bond lengths (Weber et al., 1991[Weber, E., Atwood, J. L., Davies, J. E. D. & MacNicol, D. D. (1991). Inclusion Compounds, 5th ed, pp. 188-263. Oxford University Press.], 2004[Weber, E., Korkas, P. P., Czugler, M. & Seichter, W. (2004). Supramol. Chem. 16, 217-226.]; Chandrasekhar et al., 2013[Chandrasekhar, V. & Narayanan, R. S. (2013). Indian J. Chem. Sect. A Inorg. Bio-inorg. Phys. Theor. Anal. Chem. 52, 1066-1072.]). The mol­ecular structure of (I)[link] is shown in Fig. 2[link]. The cyclo­hexane moieties of both independent mol­ecules adopt chair conformations, with atoms C1 and C4 deviating from the plane of the remaining four atoms by 0.655 and −0.657 Å, respectively, in mol­ecule A, by 0.668 and −0.638 Å in B. The disposition of the cyclo­hexane rings relative to the 1,3-diyne chain is the same in mol­ecules A and B, as shown by the similar distances C7ACg1 = 2.331 Å and C7BCg2 = 2.329 Å where Cg1 and Cg2 are the ring centroids. However, the orientation of the rings relative to each other is different (Fig. 2[link], inserts): trans in mol­ecule A, gauche in B, both different from the nearly eclipsed disposition in the one known mol­ecular complex of (I)[link].

[Figure 2]
Figure 2
The mol­ecular structure of (I)[link]. Displacement ellipsoids are drawn at the 30% probability level, hydrogen bonds are shown as dotted lines. Symmetrically independent atoms are labelled, the rest are generated by the symmetry operations 1 − x, 1 − y, 1 − z (for A) and −x, y, [{3\over 2}] − z (for B).

3. Supra­molecular features

The mol­ecule of (I)[link] has two OH groups. Each group realises its proton-donor and proton-acceptor possibilities, forming inter­molecular hydrogen bonds (Table 1[link]) O1A—H1A⋯O1B and O1B—H1B⋯O1A with O⋯O distances of 2.748 (1) and 2.771 (1) Å, respectively. As shown in Fig. 3[link], each mol­ecule participates in two [R_{4}^{4}](8) rings of hydrogen bonds (Grell et al., 1999[Grell, J., Bernstein, J. & Tinhofer, G. (1999). Acta Cryst. B55, 1030-1043.]), each ring involving two mol­ecules of type A and two of B. These bonds give rise to a two-dimensional supra­molecular layer parallel to the ac plane. The layers are incorporated into a three-dimensional network by van der Waals inter­actions (Fig. 3[link]).

Table 1
Hydrogen-bond geometry (Å, °)

D—H⋯A D—H H⋯A DA D—H⋯A
O1B—H1B⋯O1Ai 0.82 1.96 2.7709 (15) 168
O1A—H1A⋯O1B 0.82 1.94 2.7481 (15) 168
Symmetry code: (i) [-x+1, y, -z+{\script{3\over 2}}].
[Figure 3]
Figure 3
Packing diagram of (I)[link]. Dotted lines indicate hydrogen bonds. Symmetry operation for primed atoms: 1 − x, y, [{3\over 2}] − z.

4. Hirshfeld surface analysis

Hirshfeld surfaces were calculated and two-dimensional fingerprints generated using CrystalExplorer21 (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.]). Hirshfeld surfaces were obtained using a standard (high) surface resolution with the three-dimensional dnorm surfaces mapped over a fixed color scale of −0.5154 (red) to 1.9215 (blue) (Fig. 4[link]). The only red spots on the surface (revealing strong inter­actions) correspond to the O—H⋯O hydrogen bonds, the rest representing standard (white) or longer than standard (blue) van der Waals contacts. This agrees with the calculated electrostatic potential of the mol­ecule (Fig. 5[link]) where the only negative potential (acceptor) areas are around the O atoms. The two-dimensional fingerprint plots (in de vs di coordinates) (Fig. 6[link]) show that mol­ecules A and B have very similar environments, the major contributions being from contacts H⋯H (70.6 for A, 71.1% for B), H⋯C/C⋯H (18.4 and 18.7%) and H⋯O/O⋯H (11.0 and 10.2%).

[Figure 4]
Figure 4
Three-dimensional Hirshfeld surfaces of mol­ecules A and B of (I)[link] plotted over dnorm in the range −0.5154 to 1.9215 a.u.
[Figure 5]
Figure 5
Hirshfeld surfaces of mol­ecules A and B plotted over electrostatic potential in the range −0.05 to 0.05 a.u. using the B3LYP/6–311 G(d,p) basis set at the Hartree–Fock level of theory. Blue and red regions indicate positive and negative potentials, respectively.
[Figure 6]
Figure 6
Complete two-dimensional fingerprint plots for mol­ecules A (a) and B (b) of (I)[link] with relative contributions of individual contacts. Note the `spikes' indicating strong hydrogen bonds.

5. The analysis of DFT calculations

The co-presence of trans and gauche conformations of (I)[link] in the crystal was mentioned above. In order to determine the intra­molecular rotational barrier of a cyclo­hexan-1-ol fragment around the diyne rod (i.e. the Csp3—Csp bond), the relaxed scan calculation has been carried out in a vacuum by B3LYP/def2-TZVP method using the ORCA program package (Neese, 2022[Neese, F. (2022). WIREs Comput. Mol. Sci. 12, 1-15.]). The initial geometry of (I)[link] was taken from the crystal structure (CIF file) and the input files were prepared using Avogadro program package (Hanwell et al., 2012[Hanwell, M. D., Curtis, D. E., Lonie, D. C., Vandermeersch, T., Zurek, E. & Hutchison, G. R. (2012). J. Cheminform, 4, https://dx.doi.org/10.1186/1758-2946-4-17.]). The O1—C1⋯C1′—O1′ torsion angle (ω) was varied from 0 to 180° in 3° steps with full optimization of the mol­ecular geometry at each step. Then single-point calculations were performed using the B3LYP-D3BJ/def2-TZVP basis set for the geometries obtained at each step, by including dispersion corrections (Grimme et al., 2011[Grimme, S., Ehrlich, S. & Goerigk, L. (2011). J. Comput. Chem. 32, 1456-1465.]). Thus we observed energy minima at ω = 9, 30, 61, 85, 109, 146 and 180° (Fig. 7[link]), the deepest one being at 61° by DFT/def2-TZVP calculations (or 64° by DFT-D3BJ/def2-TZVP); however, the rotation barrier was low, 0.7 or 0.9 kJ mol−11, respectively. Thus, an easy transition between conformations can occur in solution and, apparently, the inter­molecular (packing) inter­actions played a decisive role in the implementation of the gauche (ω = 85°) and trans (ω = 180°) conformations in the crystal. To study the influence of ω variation on the electronic parameters, we analyzed the changes of HOMO and LUMO energies, and the energy gap upon varying ω from 0 to 180°. The energy and electron density at these orbitals are important in defining the mol­ecule's chemistry (Fukui, 1982[Fukui, K. (1982). Angew. Chem. Int. Ed. Engl. 21, 801-809.]; Hoffmann et al., 1965[Hoffmann, R. & Woodward, R. B. (1965). J. Am. Chem. Soc. 87, 2046-2048.]), the HOMO correlating with the ionization potential and representing the electron-donating ability of a mol­ecule, while the LUMO correlates with the electron affinity of a mol­ecule and represents its electron-accepting ability. The energy difference (energy gap) between HOMO and LUMO is known to represent the stability or reactivity of a mol­ecule in a series of related compounds (Pearson, 1988[Pearson, R. G. (1988). J. Am. Chem. Soc. 110, 2092-2097.]; Jahnke et al., 2010[Jahnke, E. & Tykwinski, R. R. (2010). Chem. Commun. 46, 3235-3249.]). For (I)[link], the HOMO and LUMO energies and the energy gap change slightly with ω, the former varying from −6.63 to −6.72 eV and the latter from −0.69 to −0.84 eV, while the energy gap varies from 5.79 to 5.99 eV (Fig. 8[link]). The widest energy gap (5.99 eV) was found at energetically optimal conformation with ω = 61 or 64° (vide supra). Mol­ecule (I)[link] has a low-lying HOMO and a high-lying LUMO and consequently a wide HOMO–LUMO gap, which indicates the high thermodynamic stability and low reactivity of the mol­ecule. Despite this, the highly unsaturated carbon chains could also exhibit various reaction properties (photoisomerization, nucleophilic addition of alcohols, thiols and amines to the triple bond) under special conditions (Shi et al., 2014[Shi, W. & Lei, A. (2014). Tetrahedron Lett. 55, 2763-2772.]). The reactivity of (I)[link] toward nucleophiles can be inferred from the electron density on LUMO, which is predominantly the π* orbital of di­acetyl­ene C atoms (Fig. 9[link]). The HOMO is a π-type MO and is mainly delocalized along the di­acetyl­ene fragment (Fig. 9[link]). However, these atoms are unlikely to have an electron-donating ability to electrophile reagents because of the low-lying HOMO.

[Figure 7]
Figure 7
Potential energy curve for mol­ecule (I)[link] as a function of the dihedral angle ω.
[Figure 8]
Figure 8
The HOMO–LUMO energy gap of mol­ecule (I)[link] as a function of ω.
[Figure 9]
Figure 9
(a) Optimized conformation (ω = 61°) of (I)[link] and (b) electron densities on its frontier MOs by the DFT/def2-TZVP method.

Thus, theoretical calculations showed that the rotation of hexa­nol-1 fragment around the Csp3—Csp bond can pass through several conformational minima that differ in ω. However, all these conformations make a negligible difference to the total energies and the rotational barrier between them. The conformations observed in the crystal packing arose as a result of the action of inter­molecular inter­action forces.

6. Database survey

A survey of the Cambridge Structural Database (CONQUEST version 2021 3.0; Groom et al., 2016[Groom, C. R., Bruno, I. J., Lightfoot, M. P. & Ward, S. C. (2016). Acta Cryst. B72, 171-179.]) revealed 198 structures in which an OH group and any other substituents are attached to each end of a hexa-2,4-diyne rod. However, the only compound involving compound (I)[link] is its complex with 1,4-di­aza­bicyclo [2.2.2]octane (MIRJEE; Chandrasekhar et al., 2013[Chandrasekhar, V. & Narayanan, R. S. (2013). Indian J. Chem. Sect. A Inorg. Bio-inorg. Phys. Theor. Anal. Chem. 52, 1066-1072.]). The existence of this co-crystal could be expected from the propensity of `wheel-and-axle'-shaped mol­ecules to form host–guest structures.

7. Synthesis and crystallization

The dimerization process of 1-ethynyl­cyclo­hexa­nol was conducted at 298 K for 48 h, based on a catalytic system with a copper(I) chloride catalyst, tetra­chloro­methane, N1,N1,N2,N2-tetra­methyl­ethylenedi­amine as a ligand and ethanol as the solvent, following the general routine used by Tirkasheva et al. (2022[Tirkasheva, S. I., Ziyadullaev, O. E., Muzalevskiy, V. M. & Parmanov, A. B. (2022). Molbank, 2022, M1484.]) to prepare 8,13-di­methyl­cosa-9,11-diyne-8,13-diol. This yielded 1,1′-(buta-1,3-diyne-1,4-di­yl)bis­(cyclo­hexan-1-ol) (I)[link] as a brown liquid. 25 mg (0.1 mmol) of (I)[link] were dissolved in 2 ml of chloro­form in a 50 ml round-bottom flask and the solvent was removed under vacuum. After the chloro­form was completely removed, 2 ml of CH2Cl2 and 1 ml of methanol were added to the flask. Brown single crystals of the title compound suitable for X-ray diffraction analysis were grown over three days by slow evaporation of the solvent, yield 76%, m.p. 448 K. Elemental analysis for C16H22O2 (246.33): calculated C 78.01; H 9.00%; found C 77.95; H 8.94%. FTIR (ATR), cm−1: 3326 (OH).

8. Refinement

Crystal data, data collection and structure refinement details are summarized in Table 2[link]. All H atoms were positioned geometrically, C—H 0.97 Å (methyl­ene), O—H 0.82 Å (hydroxyl group) and refined as riding with Uiso(H) = 1.2Ueq(C) or 1.5Ueq(O).

Table 2
Experimental details

Crystal data
Chemical formula C16H22O2
Mr 246.33
Crystal system, space group Monoclinic, P2/c
Temperature (K) 293
a, b, c (Å) 10.4134 (2), 6.9167 (2), 20.4801 (5)
β (°) 90.308 (2)
V3) 1475.09 (6)
Z 4
Radiation type Cu Kα
μ (mm−1) 0.56
Crystal size (mm) 0.30 × 0.24 × 0.15
 
Data collection
Diffractometer XtaLAB Synergy, Single source at home/near, HyPix3000
Absorption correction Multi-scan (CrysAlis PRO; Rigaku OD, 2020[Rigaku OD (2020). CrysAlis PRO. Oxford Diffraction Ltd, Yarnton, England.])
Tmin, Tmax 0.960, 1.000
No. of measured, independent and observed [I > 2σ(I)] reflections 13939, 2866, 2161
Rint 0.038
(sin θ/λ)max−1) 0.615
 
Refinement
R[F2 > 2σ(F2)], wR(F2), S 0.044, 0.129, 1.07
No. of reflections 2866
No. of parameters 166
H-atom treatment H-atom parameters constrained
Δρmax, Δρmin (e Å−3) 0.12, −0.16
Computer programs: CrysAlis PRO (Rigaku OD, 2020[Rigaku OD (2020). CrysAlis PRO. Oxford Diffraction Ltd, Yarnton, England.]), SHELXT2018/2 (Sheldrick, 2015a[Sheldrick, G. M. (2015a). Acta Cryst. A71, 3-8.]), SHELXL2019/2 (Sheldrick, 2015b[Sheldrick, G. M. (2015b). Acta Cryst. C71, 3-8.]), 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: CrysAlis PRO 1.171.40.84a (Rigaku OD, 2020); cell refinement: CrysAlis PRO 1.171.40.84a (Rigaku OD, 2020); data reduction: CrysAlis PRO 1.171.40.84a (Rigaku OD, 2020); program(s) used to solve structure: SHELXT2018/2 (Sheldrick, 2015a); program(s) used to refine structure: SHELXL2019/2 (Sheldrick, 2015b); molecular graphics: Mercury (Macrae et al., 2020); software used to prepare material for publication: publCIF (Westrip, 2010).

1,1'-(Buta-1,3-diyne-1,4-diyl)bis(cyclohexan-1-ol) top
Crystal data top
C16H22O2F(000) = 536
Mr = 246.33Dx = 1.109 Mg m3
Monoclinic, P2/cCu Kα radiation, λ = 1.54184 Å
a = 10.4134 (2) ÅCell parameters from 4018 reflections
b = 6.9167 (2) Åθ = 4.2–70.2°
c = 20.4801 (5) ŵ = 0.56 mm1
β = 90.308 (2)°T = 293 K
V = 1475.09 (6) Å3Block, colorless
Z = 40.30 × 0.24 × 0.15 mm
Data collection top
XtaLAB Synergy, Single source at home/near, HyPix3000
diffractometer
2866 independent reflections
Radiation source: micro-focus sealed X-ray tube, PhotonJet (Cu) X-ray Source2161 reflections with I > 2σ(I)
Mirror monochromatorRint = 0.038
Detector resolution: 10.0000 pixels mm-1θmax = 71.5°, θmin = 4.3°
ω scansh = 1212
Absorption correction: multi-scan
(CrysAlisPro; Rigaku OD, 2020)
k = 88
Tmin = 0.960, Tmax = 1.000l = 2525
13939 measured reflections
Refinement top
Refinement on F2Hydrogen site location: inferred from neighbouring sites
Least-squares matrix: fullH-atom parameters constrained
R[F2 > 2σ(F2)] = 0.044 w = 1/[σ2(Fo2) + (0.0565P)2 + 0.2398P]
where P = (Fo2 + 2Fc2)/3
wR(F2) = 0.129(Δ/σ)max = 0.001
S = 1.07Δρmax = 0.12 e Å3
2866 reflectionsΔρmin = 0.16 e Å3
166 parametersExtinction correction: SHELXL2019/2 (Sheldrick, 2015b), Fc*=kFc[1+0.001xFc2λ3/sin(2θ)]-1/4
0 restraintsExtinction coefficient: 0.0018 (4)
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
O1A0.61299 (10)0.28504 (18)0.67530 (5)0.0500 (3)
H1A0.5346510.2970810.6754850.075*
O1B0.35098 (10)0.28922 (18)0.69056 (5)0.0514 (3)
H1B0.3527760.2757540.7303310.077*
C8B0.05502 (14)0.1260 (2)0.73133 (7)0.0436 (4)
C7B0.15086 (14)0.1312 (2)0.69953 (7)0.0428 (4)
C7A0.59587 (14)0.4643 (3)0.57736 (7)0.0438 (4)
C1A0.66899 (13)0.4401 (2)0.63846 (7)0.0396 (4)
C1B0.27132 (13)0.1416 (2)0.66200 (7)0.0393 (4)
C8A0.53529 (14)0.4875 (2)0.52842 (7)0.0454 (4)
C6B0.24442 (16)0.2025 (3)0.59194 (7)0.0515 (4)
H6BA0.3250690.2261050.5698370.062*
H6BB0.1958400.3222060.5918710.062*
C2A0.80687 (15)0.3812 (3)0.62439 (9)0.0572 (5)
H2AA0.8068690.2685480.5961910.069*
H2AB0.8493390.3462810.6649710.069*
C6A0.66516 (19)0.6273 (3)0.67772 (8)0.0578 (5)
H6AA0.5765030.6662160.6836560.069*
H6AB0.7024450.6048320.7205570.069*
C2B0.33965 (16)0.0527 (3)0.66322 (10)0.0581 (5)
H2BA0.4235140.0389740.6433670.070*
H2BB0.3523510.0926770.7081920.070*
C5B0.1692 (2)0.0480 (4)0.55515 (9)0.0738 (6)
H5BA0.0842370.0362830.5739300.089*
H5BB0.1591520.0869460.5099060.089*
C3B0.2648 (2)0.2066 (3)0.62734 (12)0.0791 (7)
H3BA0.1848760.2309670.6500260.095*
H3BB0.3138850.3257360.6270150.095*
C3A0.88123 (18)0.5442 (4)0.59165 (11)0.0813 (7)
H3AA0.8459710.5672680.5484090.098*
H3AB0.9703240.5061430.5868230.098*
C5A0.7378 (2)0.7890 (3)0.64418 (11)0.0819 (7)
H5AA0.6958580.8204260.6031320.098*
H5AB0.7367690.9035810.6714910.098*
C4A0.8742 (3)0.7302 (4)0.63159 (13)0.0994 (9)
H4AA0.9183220.7112580.6729390.119*
H4AB0.9177280.8330080.6082940.119*
C4B0.2359 (3)0.1460 (4)0.55798 (13)0.0955 (9)
H4BA0.1817030.2427720.5374110.115*
H4BB0.3154490.1390780.5336870.115*
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
O1A0.0444 (6)0.0596 (7)0.0460 (6)0.0077 (5)0.0005 (5)0.0128 (5)
O1B0.0458 (6)0.0646 (8)0.0439 (6)0.0126 (5)0.0007 (5)0.0067 (6)
C8B0.0370 (7)0.0557 (10)0.0381 (8)0.0001 (7)0.0016 (6)0.0007 (7)
C7B0.0391 (8)0.0511 (9)0.0382 (8)0.0013 (7)0.0020 (6)0.0006 (7)
C7A0.0400 (8)0.0532 (10)0.0381 (8)0.0031 (7)0.0021 (6)0.0029 (7)
C1A0.0371 (7)0.0488 (9)0.0327 (7)0.0037 (6)0.0029 (6)0.0055 (6)
C1B0.0336 (7)0.0472 (9)0.0372 (8)0.0037 (6)0.0038 (6)0.0016 (6)
C8A0.0410 (8)0.0567 (10)0.0382 (8)0.0043 (7)0.0039 (6)0.0053 (7)
C6B0.0540 (9)0.0611 (11)0.0394 (8)0.0133 (8)0.0027 (7)0.0063 (8)
C2A0.0415 (8)0.0768 (13)0.0533 (10)0.0059 (8)0.0012 (7)0.0159 (9)
C6A0.0735 (12)0.0561 (11)0.0438 (9)0.0029 (9)0.0048 (8)0.0029 (8)
C2B0.0479 (9)0.0556 (11)0.0708 (12)0.0078 (8)0.0113 (8)0.0058 (9)
C5B0.0848 (14)0.0970 (17)0.0396 (9)0.0355 (13)0.0045 (9)0.0000 (10)
C3B0.0827 (14)0.0484 (11)0.1065 (19)0.0005 (10)0.0212 (13)0.0140 (12)
C3A0.0423 (10)0.120 (2)0.0817 (14)0.0086 (11)0.0073 (9)0.0352 (15)
C5A0.121 (2)0.0562 (13)0.0686 (13)0.0264 (13)0.0130 (13)0.0001 (10)
C4A0.0975 (19)0.114 (2)0.0869 (17)0.0640 (17)0.0204 (14)0.0256 (16)
C4B0.112 (2)0.0905 (19)0.0837 (17)0.0338 (16)0.0247 (14)0.0436 (15)
Geometric parameters (Å, º) top
O1A—H1A0.8200C6A—H6AB0.9700
O1A—C1A1.4365 (18)C6A—C5A1.517 (3)
O1B—H1B0.8200C2B—H2BA0.9700
O1B—C1B1.4379 (18)C2B—H2BB0.9700
C8B—C8Bi1.381 (3)C2B—C3B1.508 (3)
C8B—C7B1.195 (2)C5B—H5BA0.9700
C7B—C1B1.4764 (19)C5B—H5BB0.9700
C7A—C1A1.4709 (19)C5B—C4B1.512 (4)
C7A—C8A1.192 (2)C3B—H3BA0.9700
C1A—C2A1.521 (2)C3B—H3BB0.9700
C1A—C6A1.525 (2)C3B—C4B1.510 (4)
C1B—C6B1.520 (2)C3A—H3AA0.9700
C1B—C2B1.521 (2)C3A—H3AB0.9700
C8A—C8Aii1.384 (3)C3A—C4A1.526 (4)
C6B—H6BA0.9700C5A—H5AA0.9700
C6B—H6BB0.9700C5A—H5AB0.9700
C6B—C5B1.522 (3)C5A—C4A1.501 (4)
C2A—H2AA0.9700C4A—H4AA0.9700
C2A—H2AB0.9700C4A—H4AB0.9700
C2A—C3A1.525 (3)C4B—H4BA0.9700
C6A—H6AA0.9700C4B—H4BB0.9700
C1A—O1A—H1A109.5C3B—C2B—C1B112.03 (15)
C1B—O1B—H1B109.5C3B—C2B—H2BA109.2
C7B—C8B—C8Bi178.17 (13)C3B—C2B—H2BB109.2
C8B—C7B—C1B178.07 (17)C6B—C5B—H5BA109.3
C8A—C7A—C1A178.52 (18)C6B—C5B—H5BB109.3
O1A—C1A—C7A108.79 (12)H5BA—C5B—H5BB107.9
O1A—C1A—C2A106.62 (13)C4B—C5B—C6B111.67 (18)
O1A—C1A—C6A110.19 (12)C4B—C5B—H5BA109.3
C7A—C1A—C2A110.74 (13)C4B—C5B—H5BB109.3
C7A—C1A—C6A109.69 (14)C2B—C3B—H3BA109.4
C2A—C1A—C6A110.76 (14)C2B—C3B—H3BB109.4
O1B—C1B—C7B108.22 (12)C2B—C3B—C4B111.26 (19)
O1B—C1B—C6B106.85 (12)H3BA—C3B—H3BB108.0
O1B—C1B—C2B110.61 (13)C4B—C3B—H3BA109.4
C7B—C1B—C6B110.61 (12)C4B—C3B—H3BB109.4
C7B—C1B—C2B110.33 (13)C2A—C3A—H3AA109.4
C6B—C1B—C2B110.15 (14)C2A—C3A—H3AB109.4
C7A—C8A—C8Aii179.4 (2)C2A—C3A—C4A111.19 (18)
C1B—C6B—H6BA109.4H3AA—C3A—H3AB108.0
C1B—C6B—H6BB109.4C4A—C3A—H3AA109.4
C1B—C6B—C5B111.36 (15)C4A—C3A—H3AB109.4
H6BA—C6B—H6BB108.0C6A—C5A—H5AA109.5
C5B—C6B—H6BA109.4C6A—C5A—H5AB109.5
C5B—C6B—H6BB109.4H5AA—C5A—H5AB108.1
C1A—C2A—H2AA109.3C4A—C5A—C6A110.6 (2)
C1A—C2A—H2AB109.3C4A—C5A—H5AA109.5
C1A—C2A—C3A111.56 (16)C4A—C5A—H5AB109.5
H2AA—C2A—H2AB108.0C3A—C4A—H4AA109.3
C3A—C2A—H2AA109.3C3A—C4A—H4AB109.3
C3A—C2A—H2AB109.3C5A—C4A—C3A111.62 (18)
C1A—C6A—H6AA109.2C5A—C4A—H4AA109.3
C1A—C6A—H6AB109.2C5A—C4A—H4AB109.3
H6AA—C6A—H6AB107.9H4AA—C4A—H4AB108.0
C5A—C6A—C1A111.90 (15)C5B—C4B—H4BA109.3
C5A—C6A—H6AA109.2C5B—C4B—H4BB109.3
C5A—C6A—H6AB109.2C3B—C4B—C5B111.81 (18)
C1B—C2B—H2BA109.2C3B—C4B—H4BA109.3
C1B—C2B—H2BB109.2C3B—C4B—H4BB109.3
H2BA—C2B—H2BB107.9H4BA—C4B—H4BB107.9
O1A—C1A—C2A—C3A173.86 (15)C1B—C6B—C5B—C4B54.8 (2)
O1A—C1A—C6A—C5A173.01 (15)C1B—C2B—C3B—C4B55.6 (2)
O1B—C1B—C6B—C5B175.44 (15)C6B—C1B—C2B—C3B56.09 (19)
O1B—C1B—C2B—C3B173.98 (14)C6B—C5B—C4B—C3B54.0 (2)
C7B—C1B—C6B—C5B67.0 (2)C2A—C1A—C6A—C5A55.3 (2)
C7B—C1B—C2B—C3B66.31 (19)C2A—C3A—C4A—C5A55.5 (3)
C7A—C1A—C2A—C3A67.9 (2)C6A—C1A—C2A—C3A54.0 (2)
C7A—C1A—C6A—C5A67.3 (2)C6A—C5A—C4A—C3A56.3 (3)
C1A—C2A—C3A—C4A54.2 (2)C2B—C1B—C6B—C5B55.26 (19)
C1A—C6A—C5A—C4A56.5 (2)C2B—C3B—C4B—C5B54.2 (3)
Symmetry codes: (i) x, y, z+3/2; (ii) x+1, y+1, z+1.
Hydrogen-bond geometry (Å, º) top
D—H···AD—HH···AD···AD—H···A
O1B—H1B···O1Aiii0.821.962.7709 (15)168
O1A—H1A···O1B0.821.942.7481 (15)168
Symmetry code: (iii) x+1, y, z+3/2.
 

References

First citationAnkisetty, S. & Slattery, M. (2012). Mar. Drugs, 10, 1037–1043.  Web of Science CrossRef CAS PubMed Google Scholar
First citationArdila-Fierro, K. J., Bolm, C. & Hernández, J. G. (2019). Angew. Chem. Int. Ed. 58, 12945–12949.  CAS Google Scholar
First citationBrücner, R. (2010). Organic Mechanisms: Reactions, Stereochemistry and Synthesis, 3rd ed, pp. 6–8. Berlin, Heidelberg: Springer-Verlag.  Google Scholar
First citationCadierno, V. (2022). Catalysts, 12, 2–48.  Web of Science CrossRef Google Scholar
First citationChandrasekhar, V. & Narayanan, R. S. (2013). Indian J. Chem. Sect. A Inorg. Bio-inorg. Phys. Theor. Anal. Chem. 52, 1066–1072.  Google Scholar
First citationFukui, K. (1982). Angew. Chem. Int. Ed. Engl. 21, 801–809.  CrossRef Web of Science Google Scholar
First citationGeng, C.-A., Huang, X.-Y., Chen, X.-L., Ma, Y.-B., Rong, G.-Q., Zhao, Y., Zhang, X.-M. & Chen, J.-J. (2015). J. Ethnopharmacol. 176, 109–117.  Web of Science CrossRef CAS PubMed Google Scholar
First citationGrell, J., Bernstein, J. & Tinhofer, G. (1999). Acta Cryst. B55, 1030–1043.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationGrimme, S., Ehrlich, S. & Goerigk, L. (2011). J. Comput. Chem. 32, 1456–1465.  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 citationHanwell, M. D., Curtis, D. E., Lonie, D. C., Vandermeersch, T., Zurek, E. & Hutchison, G. R. (2012). J. Cheminform, 4, https://dx.doi.org/10.1186/1758-2946-4-17Web of Science CrossRef Google Scholar
First citationHoffmann, R. & Woodward, R. B. (1965). J. Am. Chem. Soc. 87, 2046–2048.  CrossRef CAS Web of Science Google Scholar
First citationHosseini, A. & Schreiner, P. R. (2020). Eur. J. Org. Chem. pp. 4339–4346.  Web of Science CrossRef Google Scholar
First citationJahnke, E. & Tykwinski, R. R. (2010). Chem. Commun. 46, 3235–3249.  Web of Science CSD CrossRef CAS Google Scholar
First citationKuang, Z., Chen, H., Yan, J., Yang, K., Lan, Y. & Song, Q. (2018). Org. Lett. 20, 5153–5157.  Web of Science CrossRef CAS PubMed Google Scholar
First citationLee, C.-Y., Yun, J. H., Kang, K., Nho, C.-W. & Shin, D. (2015). Bioorg. Med. Chem. Lett. 25, 4020–4023.  Web of Science CrossRef CAS PubMed 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 citationMotoki, R., Kanai, M. & Shibasaki, M. (2007). Org. Lett. 9, 2997–3000.  Web of Science CrossRef PubMed CAS Google Scholar
First citationNeese, F. (2022). WIREs Comput. Mol. Sci. 12, 1–15.  Web of Science CrossRef Google Scholar
First citationPearson, R. G. (1988). J. Am. Chem. Soc. 110, 2092–2097.  CrossRef CAS Web of Science Google Scholar
First citationRigaku OD (2020). CrysAlis PRO. Oxford Diffraction Ltd, Yarnton, England.  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 citationShi, W. & Lei, A. (2014). Tetrahedron Lett. 55, 2763–2772.  Web of Science CrossRef CAS Google Scholar
First citationShi Shun, A. L. K. & Tykwinski, R. R. (2006). Angew. Chem. Int. Ed. 45, 1034–1057.  Web of Science CrossRef 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 citationSum, Y. N., Yu, D. & Zhang, Y. (2013). Green Chem. 15, 2718–2721.  Web of Science CrossRef CAS Google Scholar
First citationTanaka, K., Kukita, K., Ichibakase, T., Kotani, S. & Nakajima, M. (2011). Chem. Commun. 47, 5614–5616.  Web of Science CrossRef CAS Google Scholar
First citationTirkasheva, S. I., Ziyadullaev, O. E., Muzalevskiy, V. M. & Parmanov, A. B. (2022). Molbank, 2022, M1484.  Web of Science CrossRef Google Scholar
First citationWang, Y.-H., Liu, H., Zhu, L.-L., Li, X.-X. & Chen, Z. (2011). Adv. Synth. Catal. 353, 707–712.  Web of Science CrossRef CAS Google Scholar
First citationWeber, E., Atwood, J. L., Davies, J. E. D. & MacNicol, D. D. (1991). Inclusion Compounds, 5th ed, pp. 188–263. Oxford University Press.  Google Scholar
First citationWeber, E., Korkas, P. P., Czugler, M. & Seichter, W. (2004). Supramol. Chem. 16, 217–226.  Web of Science CSD CrossRef CAS Google Scholar
First citationWestrip, S. P. (2010). J. Appl. Cryst. 43, 920–925.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationZhang, X., Teo, W. T., Sally & Chan, P. W. H. (2010). J. Org. Chem. 75, 6290–6293.  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