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

Journal logoCRYSTALLOGRAPHIC
COMMUNICATIONS
ISSN: 2056-9890

2-Methyl-4-(4-nitro­phen­yl)but-3-yn-2-ol: crystal structure, Hirshfeld surface analysis and computational chemistry study

CROSSMARK_Color_square_no_text.svg

aDepartamento de Física, Universidade Federal de São Carlos, 13565-905 São Carlos, SP, Brazil, bDepartamento de Química, Universidade Federal de São Carlos, 13565-905 São Carlos, SP, Brazil, cDepartment of Physics, Bhavan's Sheth R. A. College of Science, Ahmedabad, Gujarat 380001, India, and dResearch Centre for Crystalline Materials, School of Science and Technology, Sunway University, 47500 Bandar Sunway, Selangor Darul Ehsan, Malaysia
*Correspondence e-mail: edwardt@sunway.edu.my

Edited by W. T. A. Harrison, University of Aberdeen, Scotland (Received 17 July 2019; accepted 18 July 2019; online 23 July 2019)

The di-substituted acetyl­ene residue in the title compound, C11H11NO3, is capped at either end by di-methyl­hydroxy and 4-nitro­benzene groups; the nitro substituent is close to co-planar with the ring to which it is attached [dihedral angle = 9.4 (3)°]. The most prominent feature of the mol­ecular packing is the formation, via hy­droxy-O—H⋯O(hy­droxy) hydrogen bonds, of hexa­meric clusters about a site of symmetry [\overline{3}]. The aggregates are sustained by 12-membered {⋯OH}6 synthons and have the shape of a flattened chair. The clusters are connected into a three-dimensional architecture by benzene-C—H⋯O(nitro) inter­actions, involving both nitro-O atoms. The aforementioned inter­actions are readily identified in the calculated Hirshfeld surface. Computational chemistry indicates there is a significant energy, primarily electrostatic in nature, associated with the hy­droxy-O—H⋯O(hy­droxy) hydrogen bonds. Dispersion forces are more important in the other identified but, weaker inter­molecular contacts.

1. Chemical context

Protected acetyl­enes represent a highly privileged class of synthetic inter­mediates for the construction of a variety of different organic compounds (Tan et al., 2013[Tan, X., Kong, L., Dai, H., Cheng, X., Liu, F. & Tschierske, C. (2013). Chem. Eur. J. 19, 16303-16313.]). The preparation of protected aryl­acetyl­enes can be achieved by the palladium-catalysed Sonogashira cross-coupling of mono-protected acetyl­enes, such as tri­methyl­silyl­acetyl­ene (TMSA), triisopropysilyl­acetyl­ene (TIPSA) and 2-methyl-3-butyn-2-ol (MEBYNOL), with aryl halides (Hundertmark et al., 2000[Hundertmark, T., Littke, A. F., Buchwald, S. L. & Fu, G. C. (2000). Org. Lett. 2, 1729-1731.]; Erdélyi & Gogoll, 2001[Erdélyi, M. & Gogoll, A. (2001). J. Org. Chem. 66, 4165-4169.]). Despite the relevance of protected acetyl­enes, the release of the protecting group remains a challenge. While tri­alkyl­silyl groups can be readily removed by treatment with bases or fluoride salts under mild reaction conditions, tri­alkyl­silyl­acetyl­enes are rather expensive, in comparison to MEYBNOL, thereby limiting their use to small-scale synthesis. Thus, MEBYNOL can be viewed as one alternative to other acetyl­ene sources. Nevertheless, the reaction conditions for the release of the 2-hy­droxy­isopropyl protecting group usually requires harsh reaction conditions. Hence, several synthetic routes combine the release of the terminal acetyl­ene with a further transformation, without the isolation of the inter­mediate (Li et al., 2015[Li, X., Sun, S., Yang, F., Kang, J., Wu, Y. & Wu, Y. (2015). Org. Biomol. Chem. 13, 2432-2436.]). It was in the context of such considerations that the title acetyl­ene compound, (I)[link], previously reported (Bleicher et al., 1998[Bleicher, L. S., Cosford, N. D. P., Herbaut, A., McCallum, J. S. & McDonald, I. A. (1998). J. Org. Chem. 63, 1109-1118.]), was isolated and crystallized. Herein, the crystal and mol­ecular structures of (I)[link] are described along with a detailed analysis of the mol­ecular packing by Hirshfeld surface analysis, non-covalent inter­action plots and computational chemistry.

[Scheme 1]

2. Structural commentary

The mol­ecular structure of (I)[link], Fig. 1[link], features a di-substituted acetyl­ene residue. At one end, the acetyl­ene terminates with a di-methyl­hydroxy substituent and at the other end, with a 4-nitro­benzene group. The nitro group is slightly inclined out of the plane of the benzene ring to which it is connected, with the dihedral angle between the planes being 9.4 (3)°.

[Figure 1]
Figure 1
The mol­ecular structure of (I)[link], showing the atom-labelling scheme and displacement ellipsoids at the 35% probability level.

3. Supra­molecular features

The spectacular feature of the mol­ecular packing of (I)[link] is the presence of hexa­meric clusters connected by hy­droxy-O—H⋯O(hy­droxy) hydrogen bonds, Table 1[link]. As seen from Fig. 2[link](a), the six-mol­ecule aggregates are sustained by 12-membered {⋯OH}6 synthons. The aggregates are disposed about a site of symmetry [\overline{3}] so the rings have the shape of a flattened chair, Fig. 2[link](b). The crystal also features weak benzene-C—H⋯O(nitro) inter­actions, involving both nitro-O atoms. In essence, one nitro group of one mol­ecule forms two such inter­actions with two symmetry-related mol­ecules to form a supra­molecular chain along the c-axis direction with helical symmetry (31 screw axis), Fig. 3[link](a). An end-on view of the chain is shown in Fig. 3[link](b). These weak benzene-C—H⋯O(nitro) inter­actions serve to link the six-mol­ecule aggregates into a three-dimensional architecture, Fig. 4[link].

Table 1
Hydrogen-bond geometry (Å, °)

D—H⋯A D—H H⋯A DA D—H⋯A
O1—H1O⋯O1i 0.82 1.87 2.682 (2) 173
C10—H10⋯O3ii 0.93 2.67 3.548 (3) 157
C11—H11⋯O2iii 0.93 2.68 3.467 (3) 143
Symmetry codes: (i) [x-y+{\script{1\over 3}}, x-{\script{1\over 3}}, -z+{\script{5\over 3}}]; (ii) [-x+y+{\script{2\over 3}}, -x+{\script{4\over 3}}, z+{\script{1\over 3}}]; (iii) [-y+{\script{4\over 3}}, x-y+{\script{2\over 3}}, z+{\script{2\over 3}}].
[Figure 2]
Figure 2
Hydrogen bonding in the crystal of (I)[link]: (a) an end-on view of the hexa­gon sustained by hy­droxy-O—H⋯O(hy­droxy) hydrogen bonding (shown as orange dashed lines) and (b) a side-on view. Non-participating hydrogen atoms have been removed for reasons of clarity.
[Figure 3]
Figure 3
Details of benzene-C—H⋯O(nitro) inter­actions (shown as blue dashed lines) in the crystal of (I)[link]: (a) a view of the supra­molecular chain along the c-axis direction and (b) an end-on view of the chain.
[Figure 4]
Figure 4
A view of the unit-cell contents of (I)[link] shown in projection down the c axis. The hy­droxy-O—H⋯O(hy­droxy) hydrogen bonding and benzene-C—H⋯O(nitro) inter­actions are shown as orange and blue dashed lines, respectively.

4. Hirshfeld surface analysis

The Hirshfeld surface calculations for (I)[link] were performed in accord with protocols described in a recently published paper (Tan et al., 2019[Tan, S. L., Jotani, M. M. & Tiekink, E. R. T. (2019). Acta Cryst. E75, 308-318.]) employing Crystal Explorer 17 (Turner et al., 2017[Turner, M. J., Mckinnon, J. J., Wolff, S. K., Grimwood, D. J., Spackman, P. R., Jayatilaka, D. & Spackman, M. A. (2017). Crystal Explorer 17. The University of Western Australia.]). On the Hirshfeld surfaces mapped over dnorm in Fig. 5[link](a), the donors and acceptors of O—H⋯O hydrogen bond involving the atoms of the hydroxyl group are characterized as bright-red spots. The faint-red spots near the phenyl-H10, H11 and nitro-O2, O3 atoms on the dnorm-mapped Hirshfeld surface in Fig. 5[link](b) represent the effect of weak C—H⋯O inter­actions as listed in Table 1[link]. The Hirshfeld surface mapped over electrostatic potential in Fig. 6[link] also illustrates the donors and acceptors of the indicated inter­actions through blue and red regions corresponding to positive and negative electrostatic potentials, respectively. In the view of a surface mapped with the shape-index property, Fig. 7[link](a), the C—H⋯π/π⋯H—C contacts listed in Table 2[link] are evident as the blue bump and a bright-orange region about the participating atoms. The overlap between benzene (C6–C11) ring of a reference mol­ecule within the Hirshfeld surface mapped over curvedness and the symmetry related ring, Fig. 7[link](b) is an indication of the ππ stacking inter­action between them [centroid–centroid distance = 3.7873 (14) Å; symmetry operation: 1 − x, 1 − y, 1 − z].

Table 2
Summary of short inter­atomic contacts (Å) in (I)

The inter­atomic distances are calculated in Crystal Explorer 17 (Turner et al., 2017[Turner, M. J., Mckinnon, J. J., Wolff, S. K., Grimwood, D. J., Spackman, P. R., Jayatilaka, D. & Spackman, M. A. (2017). Crystal Explorer 17. The University of Western Australia.]) whereby the X—H bond lengths are adjusted to their neutron values.

Contact Distance Symmetry operation
O1⋯H3A 2.71 [{1\over 3}] + y, [{2\over 3}] − x + y, [{5\over 3}] − z
O2⋯H2B 2.69 [{2\over 3}] − y, [{1\over 3}] + x − y, −[{2\over 3}] + z
O3⋯H2A 2.69 1 − x, 1 − y, 1 − z
C1⋯H1O 2.85 [{1\over 3}] + y, [{2\over 3}] − x + y, [{5\over 3}] − z
C5⋯H3C 2.79 [{1\over 3}] + y, [{2\over 3}] − x + y, [{2\over 3}] − z
C7⋯H2C 2.85 [{1\over 3}] + y, [{2\over 3}] − x + y, [{2\over 3}] − z
C8⋯H2C 2.80 [{1\over 3}] + y, [{2\over 3}] − x + y, [{2\over 3}] − z
[Figure 5]
Figure 5
Two views of the Hirshfeld surface for (I)[link] mapped over dnorm: (a) in the range −0.202 to +1.400 arbitrary units and (b) in the range −0.102 to +1.400 arbitrary units, highlighting, respectively, inter­molecular O—H⋯O and C—H⋯O inter­actions through black dashed lines.
[Figure 6]
Figure 6
A view of the Hirshfeld surface for (I)[link] mapped over the electrostatic potential in the range −0.098 to + 0.180 atomic units. The red and blue regions represent negative and positive electrostatic potentials, respectively, and show the acceptors and donors of inter­molecular inter­actions, respectively.
[Figure 7]
Figure 7
(a) A view of the Hirshfeld surface for (I)[link] mapped with the shape-index property, highlighting inter­molecular C—H⋯π/π⋯H—C contacts by blue bumps and bright-orange concave regions, respectively, and (b) a view of the Hirshfeld surface mapped over curvedness, highlighting ππ contacts between symmetry-related (C6-C11) rings.

The overall two-dimensional fingerprint plot for (I)[link], Fig. 8[link](a), and those delineated into H⋯H, O⋯H/H⋯O, C⋯H/H⋯C and C⋯C contacts (McKinnon et al., 2007[McKinnon, J. J., Jayatilaka, D. & Spackman, M. A. (2007). Chem. Commun. pp. 3814-3816.]) are illustrated in Fig. 8[link](b)–(e), respectively, and provide more information on the influence of short inter­atomic contacts upon the mol­ecular packing. The percentage contributions from the different inter­atomic contacts to the Hirshfeld surface are summarized in Table 3[link]. The greatest contribution to the Hirshfeld surface of 38.2% are derived from H⋯H contacts but these exert a negligible influence on the packing, at least in terms of directional inter­actions, as the inter­atomic distances are greater than sum of their van der Waals radii. The pair of long spikes with their tips at de + di ∼1.8 Å in the fingerprint plot delineated into O⋯H/H⋯O contacts, Fig. 8[link](c), are due to the presence of the O—H⋯O hydrogen bond, whereas the points corresponding to comparatively weak inter­molecular C—H⋯O inter­actions, Table 1[link], and the short inter­atomic O⋯H/H⋯O contacts are merged within the plot, Table 2[link]. The presence of the C—H⋯π contact, formed by the methyl-H2C atom and the benzene (C6–C11) ring, results in short inter­atomic C⋯H/H⋯C contacts, Table 2[link] and Fig. 7[link](a), and by the pair of forceps-like tips at de + di ∼2.8 Å in Fig. 8[link](d). The points corresponding to other such short inter­atomic contacts involving the acetyl­ene-C5 and methyl-C3—H3c atoms at longer separations are merged within the plot. The arrow-shaped distribution of points around de + di ∼3.6 Å in the fingerprint plot delineated into C⋯C contacts, Fig. 8[link](e), indicate ππ overlap between symmetry-related benzene (C6–C11) rings, as illustrated in Fig. 7[link](b). The small percentage contributions from the other inter­atomic contacts listed in Table 3[link] have negligible influence upon the mol­ecular packing as their separations are greater than the sum of the respective van der Waals radii.

Table 3
Percentage contributions of inter­atomic contacts to the Hirshfeld surface for (I)

Contact Percentage contribution
H⋯H 38.2
O⋯H/H⋯O 32.1
C⋯H/H⋯C 20.0
C⋯C 4.2
N⋯O/O⋯N 1.7
O⋯O 1.6
C⋯N/N⋯C 1.0
N⋯H/H⋯N 0.8
C⋯O/O⋯C 0.4
[Figure 8]
Figure 8
(a) The full two-dimensional fingerprint plot for (I)[link] and (b)–(e) those delineated into H⋯H, O⋯H/H⋯O, C⋯H/H⋯C and C⋯C contacts, respectively.

5. Inter­action energies

The pairwise inter­action energies between the mol­ecules within the crystal were calculated by summing up four energy components comprising electrostatic (Eele), polarization (Epol), dispersion (Edis) and exchange-repulsion (Erep) terms after applying relevant scale factors (Turner et al., 2017[Turner, M. J., Mckinnon, J. J., Wolff, S. K., Grimwood, D. J., Spackman, P. R., Jayatilaka, D. & Spackman, M. A. (2017). Crystal Explorer 17. The University of Western Australia.]). These energies were obtained by using the wave function calculated at the B3LYP/6-31G(d,p) level. The strength and the nature of inter­molecular inter­actions in terms of their energies are qu­anti­tatively summarized in Table 4[link]. The energies calculated for the different inter­molecular inter­actions indicate that the electrostatic contribution is dominant in the O—H⋯O hydrogen bond whereas the dispersive component has a significant influence due to the presence of short inter­atomic C⋯H/H⋯C and O⋯H/H⋯O contacts occurring between the same pair of mol­ecules. The C—H⋯O2(nitro) inter­action has almost the same contributions from the electrostatic and dispersive components. This is in contrast to a major contribution only from the dispersive component for the analogous contact involving the nitro-O3 atom. The dispersion energy component makes the major contribution to the relevant pairs of mol­ecules involved in other short inter­atomic contacts, Table 4[link], as well as in C—H⋯π and ππ stacking inter­actions. It is also evident from a comparison of the total energies of inter­molecular inter­actions, Table 4[link], that the O—H⋯O hydrogen bond and ππ stacking inter­action are stronger than the other inter­actions, and, of these, the inter­molecular C—H⋯O contacts are weaker than the C—H⋯π inter­actions.

Table 4
Summary of inter­action energies (kJ mol−1) calculated for (I)

Contact R (Å) Eele Epol Edis Erep Etot
O1—H1O⋯O1i            
H3A⋯O1i 8.80 −52.3 −12.0 −18.8 72.7 −35.7
H1O⋯C1i            
C10—H10⋯O3ii 8.28 −3.7 −1.4 −9.2 4.9 −9.8
C11—H11⋯O2iii 9.51 −5.8 −1.7 −5.7 5.0 −9.6
O3⋯H2Aiv            
(C6–C11)⋯(C6–C11)iv 4.25 −9.4 −1.8 −47.1 28.9 −34.4
H3C⋯C5v            
H2C⋯C7v            
H2C⋯C8v 5.78 −2.1 −0.7 −28.6 18.2 −16.4
C2—H2C⋯(C6–C11)v            
Symmetry codes: (i) [{1\over 3}] + x − y, [{1\over 3}] − x, [{5\over 3}] − z; (ii) [{2\over 3}] − x + y, 4/3 − x, [{1\over 3}] + z; (iii) 4/3 − y, [{2\over 3}] + x − y, [{2\over 3}] + z; (iv) 1 − x, 1 − y, 1 − z; (v) [{1\over 3}] + x − y, − [{1\over 3}] + x, [{2\over 3}] − z.

The magnitudes of inter­molecular energies are represented graphically by energy frameworks to view the supra­molecular architecture of the crystal through the cylinders joining centroids of mol­ecular pairs by using red, green and blue colour codes for the components Eele, Edisp and Etot, respectively, Fig. 9[link]. The radius of the cylinder is proportional to the magnitude of inter­action energy, which are adjusted to the same scale factor of 30 with a cut-off value of 3 kJ mol−1 within 2 × 2 × 2 unit cells.

[Figure 9]
Figure 9
A comparison of the energy frameworks calculated for (I)[link] and viewed down the c axis showing (a) electrostatic potential force, (b) dispersion force and (c) total energy. The energy frameworks were adjusted to the same scale factor of 30 with a cut-off value of 3 kJ mol−1 within 2 × 2 × 2 unit cells.

6. Non-covalent inter­action plots

Non-covalent inter­action plot (NCIplot) analyses provide a visual representation of the nature of the contact between specified species in crystals (Johnson et al., 2010[Johnson, E. R., Keinan, S., Mori-Sánchez, P., Contreras-García, J., Cohen, A. J. & Yang, W. (2010). J. Am. Chem. Soc. 132, 6498-6506.]; Contreras-Garcá et al., 2011[Contreras-García, J., Johnson, E. R., Keinan, S., Chaudret, R., Piquemal, J.-P., Beratan, D. N. & Yang, W. (2011). J. Chem. Theory Comput. 7, 625-632.]). This method is based on the electron density (and derivatives) and was employed in the present study to confirm the nature of some of the specified inter­molecular contacts. The colour-based isosurfaces generated correspond to the values of sign(λ2)ρ(r), where ρ is the electron density and λ2 is the second eigenvalue of the Hessian matrix of ρ. Crucially, through a three-colour scheme, a specific inter­action can be identified as being attractive or otherwise. Thus, a green isosurface indicates a weakly attractive inter­action whereas a blue isosurface indicates an attractive inter­action; a repulsive inter­action appears red. The isosurfaces for three identified inter­molecular inter­actions are given in the upper view of Fig. 10[link]. Thus, in Fig. 10[link](a), a green isosurface is apparent for the conventional hy­droxy-O—H⋯O(hy­droxy) hydrogen bond. Similarly, green isosurfaces are seen between the inter­acting atoms involved in the phenyl-C—H⋯O(nitro), Fig. 10[link](b), and the methyl-C—H⋯π(C11–C16), Fig. 10[link](c), inter­actions.

[Figure 10]
Figure 10
Non-covalent inter­action plots for (a) hy­droxy-O—H⋯O(hy­droxy) hydrogen bonding, (b) the phenyl-C—H⋯O(nitro) inter­actions and (c) the methyl-C—H⋯π(C11–C16) inter­actions.

The lower views of Fig. 10[link], show the plots of the RDG versus sign(λ2)ρ(r). The non-covalent inter­action peaks appear at density values less than 0.0 atomic units, consistent with their being weakly attractive inter­actions.

7. Database survey

There are four literature precedents for (I)[link] with varying substitution patterns in the appended benzene ring. These are the unsubstituted `parent' compound [(II); FESMEV; Singelenberg & van Eijck, 1987[Singelenberg, F. A. J. & van Eijck, B. P. (1987). Acta Cryst. C43, 693-695.]], and the 4-cyano [(III}; HEFDAA; Clegg, 2017[Clegg, W. (2017). Private communication (refcode: HEFDAA). CCDC, Cambridge, England.]], 4-meth­oxy [(IV); YUQPEG; Eissmann et al., 2010[Eissmann, F., Kafurke, U. & Weber, E. (2010). Acta Cryst. E66, o1866.]] and 3-acetyl-4-hy­droxy [(V); UVETAS; Hübscher et al., 2016[Hübscher, J., Rosin, R., Seichter, W. & Weber, E. (2016). Acta Cryst. E72, 1370-1373.]] derivatives. Selected geometric parameters for (I)–(IV) are collated in Table 5[link]. Of particular inter­est in the mode of supra­molecular association in their crystals. As seen from Fig. 11[link], four distinct patterns appear. In (V), three independent mol­ecules comprise the asymmetric unit and these associate about a centre of inversion in space group P21/c to form a hexa­meric clusters via hy­droxy-O—H⋯O(hy­droxy) hydrogen bonds as seen in (I)[link], Fig. 11[link](a); intra­molecular hy­droxy-O—H⋯O(carbon­yl) hydrogen bonds are also apparent. In (III), the two independent mol­ecules comprising the asymmetric unit associate about a centre of inversion in space group P21/n into a supra­molecular dimer via pairs of hy­droxy-O—H⋯O(hy­droxy) and hy­droxy-O—H⋯N(cyano) hydrogen bonds as shown in Fig. 11[link](b). In this case, one independent hy­droxy-oxygen atom and one cyano-nitro­gen atom do not accept a hydrogen-bonding inter­action. Three crystallographically independent mol­ecules are also found in (II) (space group Pca21) and these self-associate to form a supra­molecular chain via hy­droxy-O—H⋯O(hy­droxy) hydrogen bonds with non-crystallographic threefold symmetry, Fig. 11[link](c). Finally, zigzag supra­molecular chains sustained by hy­droxy-O—H⋯O(hy­droxy) hydrogen bonds are found in the crystal of (IV), Fig. 11[link](d) in space group Pbca.

Table 5
Geometric data (Å, °) for related 2-methyl-4-(ar­yl)but-3-yn-2-ol mol­ecules

Compound Z Cring—Cacetyl­ene Cacetyl­ene—Cacetyl­ene Cacetyl­ene—Cquaternary Supra­molecular motif Reference
(I) 1 1.438 (3) 1.189 (3) 1.471 (3) hexa­mer This work
(II) 3 1.443 (5) 1.211 (5) 1.454 (5) chain Singelenberg & van Eijck (1987[Singelenberg, F. A. J. & van Eijck, B. P. (1987). Acta Cryst. C43, 693-695.])
    1.437 (6) 1.192 (6) 1.479 (6)    
    1.437 (5) 1.189 (5) 1.479 (5)    
(III) 2 1.441 (2) 1.193 (2) 1.490 (2) dimer Clegg (2017[Clegg, W. (2017). Private communication (refcode: HEFDAA). CCDC, Cambridge, England.])
    1.435 (2) 1.1895 (2) 1.480 (2)    
(IV) 1 1.4377 (16) 1.2000 (16) 1.4791 (16) chain Eissmann et al. (2010[Eissmann, F., Kafurke, U. & Weber, E. (2010). Acta Cryst. E66, o1866.])
(V) 3 1.4418 (18) 1.1951 (19) 1.4764 (19) hexa­mer Hübscher et al. (2016[Hübscher, J., Rosin, R., Seichter, W. & Weber, E. (2016). Acta Cryst. E72, 1370-1373.])
    1.444 (2) 1.194 (2) 1.4859 (19)    
    1.4402 (19) 1.1904 (19) 1.4723 (18)    
[Figure 11]
Figure 11
Supra­molecular association via hy­droxy-O—H⋯O(hy­droxy) hydrogen bonds in (II)–(IV): (a) hexa­meric cluster in (V), (b) dimeric aggregate sustained by additional hy­droxy-O—H⋯N(cyano) hydrogen bonds in (III), (c) views of the supra­molecular chain in (II) with non-crystallographic threefold symmetry and (d) views of the zigzag supra­molecular chain in (IV).

8. Synthesis and crystallization

The title compound was prepared as per the literature procedure (Bleicher et al., 1998[Bleicher, L. S., Cosford, N. D. P., Herbaut, A., McCallum, J. S. & McDonald, I. A. (1998). J. Org. Chem. 63, 1109-1118.]). Yield: 87%. Yellow solid, m.p. 377–379 K. 1H NMR (400 MHz, CDCl3): δ = 8.16 (dt, J = 8.9, 2.2 Hz, 2H), 7.54 (dt, J = 8.9, 2.2 Hz, 2H), 2.24 (s, 1H) and 1.63 (s, 6H) ppm. 13C NMR (101 MHz, CDCl3): δ = 147.2, 132.5, 129.8, 123.6, 99.2, 80.5, 66.7 and 31.3 ppm. Irregular colourless crystals of (I)[link] for the X-ray study were grown by slow evaporation of its ethyl acetate solution.

9. Refinement details

Crystal data, data collection and structure refinement details are summarized in Table 6[link]. The carbon-bound H atoms were placed in calculated positions (C—H = 0.93–0.96 Å) and were included in the refinement in the riding-model approximation, with Uiso(H) set to 1.2–1.5Ueq(C). The O-bound H atom was refined with a distance restraint of 0.82±0.01 Å, and with Uiso(H) = 1.5Ueq(O).

Table 6
Experimental details

Crystal data
Chemical formula C11H11NO3
Mr 205.21
Crystal system, space group Trigonal, R[\overline{3}]:H
Temperature (K) 296
a, c (Å) 26.3146 (14), 8.1205 (5)
V3) 4869.8 (6)
Z 18
Radiation type Mo Kα
μ (mm−1) 0.09
Crystal size (mm) 0.34 × 0.28 × 0.16
 
Data collection
Diffractometer Bruker APEXII CCD
Absorption correction Multi-scan (SADABS; Sheldrick, 1996[Sheldrick, G. M. (1996). SADABS. University of Göttingen, Germany.])
Tmin, Tmax 0.440, 0.745
No. of measured, independent and observed [I > 2σ(I)] reflections 10643, 2230, 1513
Rint 0.080
(sin θ/λ)max−1) 0.627
 
Refinement
R[F2 > 2σ(F2)], wR(F2), S 0.053, 0.149, 1.05
No. of reflections 2230
No. of parameters 139
No. of restraints 1
H-atom treatment H-atom parameters constrained
Δρmax, Δρmin (e Å−3) 0.16, −0.27
Computer programs: APEX2 and SAINT (Bruker, 2009[Bruker (2009). APEX2 and SAINT. Bruker AXS Inc., Madison, Wisconsin, USA.]), SIR2014 (Burla et al., 2015[Burla, M. C., Caliandro, R., Carrozzini, B., Cascarano, G. L., Cuocci, C., Giacovazzo, C., Mallamo, M., Mazzone, A. & Polidori, G. (2015). J. Appl. Cryst. 48, 306-309.]), SHELXL2018/3 (Sheldrick, 2015[Sheldrick, G. M. (2015). Acta Cryst. C71, 3-8.]), ORTEP-3 for Windows (Farrugia, 2012[Farrugia, L. J. (2012). J. Appl. Cryst. 45, 849-854.]), DIAMOND (Brandenburg, 2006[Brandenburg, K. (2006). DIAMOND. Crystal Impact GbR, Bonn, Germany.]), MarvinSketch (ChemAxon, 2010[ChemAxon (2010). Marvinsketch. https://www.chemaxon.com.]) and publCIF (Westrip, 2010[Westrip, S. P. (2010). J. Appl. Cryst. 43, 920-925.]).

Supporting information


Computing details top

Data collection: APEX2 (Bruker, 2009); cell refinement: SAINT (Bruker, 2009); data reduction: SAINT (Bruker, 2009); program(s) used to solve structure: SIR2014 (Burla et al., 2015); program(s) used to refine structure: SHELXL2018/3 (Sheldrick, 2015); molecular graphics: ORTEP-3 for Windows (Farrugia, 2012), DIAMOND (Brandenburg, 2006); software used to prepare material for publication: MarvinSketch (ChemAxon, 2010) and publCIF (Westrip, 2010).

2-Methyl-4-(4-nitrophenyl)but-3-yn-2-ol top
Crystal data top
C11H11NO3Dx = 1.260 Mg m3
Mr = 205.21Mo Kα radiation, λ = 0.71073 Å
Trigonal, R3:HCell parameters from 2006 reflections
a = 26.3146 (14) Åθ = 2.7–23.9°
c = 8.1205 (5) ŵ = 0.09 mm1
V = 4869.8 (6) Å3T = 296 K
Z = 18Irregular, colourles
F(000) = 19440.34 × 0.28 × 0.16 mm
Data collection top
Bruker APEXII CCD
diffractometer
1513 reflections with I > 2σ(I)
φ and ω scansRint = 0.080
Absorption correction: multi-scan
(SADABS; Sheldrick, 1996)
θmax = 26.4°, θmin = 1.6°
Tmin = 0.440, Tmax = 0.745h = 3232
10643 measured reflectionsk = 3232
2230 independent reflectionsl = 910
Refinement top
Refinement on F2Primary atom site location: structure-invariant direct methods
Least-squares matrix: fullSecondary atom site location: difference Fourier map
R[F2 > 2σ(F2)] = 0.053Hydrogen site location: inferred from neighbouring sites
wR(F2) = 0.149H-atom parameters constrained
S = 1.05 w = 1/[σ2(Fo2) + (0.0511P)2 + 3.9317P]
where P = (Fo2 + 2Fc2)/3
2230 reflections(Δ/σ)max < 0.001
139 parametersΔρmax = 0.16 e Å3
1 restraintΔρmin = 0.27 e Å3
Special details top

Geometry. All esds (except the esd in the dihedral angle between two l.s. planes) are estimated using the full covariance matrix. The cell esds are taken into account individually in the estimation of esds in distances, angles and torsion angles; correlations between esds in cell parameters are only used when they are defined by crystal symmetry. An approximate (isotropic) treatment of cell esds is used for estimating esds involving l.s. planes.

Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/Ueq
O10.56993 (7)0.33529 (6)0.78650 (17)0.0513 (4)
H1O0.5692410.3059730.8236590.077*
O20.54778 (9)0.61319 (8)0.0578 (2)0.0811 (6)
O30.62154 (8)0.66736 (8)0.2126 (3)0.0785 (6)
N10.58062 (9)0.61978 (9)0.1723 (3)0.0565 (5)
C10.53437 (9)0.31996 (8)0.6425 (2)0.0388 (5)
C20.47123 (10)0.27726 (10)0.6889 (3)0.0612 (7)
H2A0.4591500.2938920.7751020.092*
H2B0.4680580.2412000.7266200.092*
H2C0.4465100.2696370.5944580.092*
C30.55665 (12)0.29361 (11)0.5140 (3)0.0634 (7)
H3A0.5516380.2570440.5539540.095*
H3B0.5975040.3202350.4932790.095*
H3C0.5348600.2869510.4137690.095*
C40.54001 (9)0.37468 (9)0.5772 (2)0.0439 (5)
C50.54464 (10)0.41762 (9)0.5145 (2)0.0458 (5)
C60.55281 (9)0.46950 (9)0.4317 (2)0.0407 (5)
C70.51138 (9)0.46599 (9)0.3192 (2)0.0424 (5)
H70.4775590.4301840.3002980.051*
C80.52018 (9)0.51543 (9)0.2351 (2)0.0441 (5)
H80.4925230.5132840.1599330.053*
C90.57045 (9)0.56768 (9)0.2648 (2)0.0416 (5)
C100.61185 (10)0.57274 (9)0.3773 (3)0.0515 (6)
H100.6452620.6088210.3968420.062*
C110.60276 (10)0.52332 (10)0.4602 (3)0.0509 (6)
H110.6304220.5260000.5362420.061*
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
O10.0636 (10)0.0378 (8)0.0474 (8)0.0216 (7)0.0238 (7)0.0036 (6)
O20.0870 (14)0.0812 (13)0.0791 (12)0.0450 (11)0.0075 (11)0.0315 (10)
O30.0734 (13)0.0474 (10)0.1082 (15)0.0254 (10)0.0056 (11)0.0178 (10)
N10.0592 (12)0.0508 (12)0.0661 (13)0.0324 (11)0.0118 (10)0.0164 (9)
C10.0463 (11)0.0376 (10)0.0322 (9)0.0206 (9)0.0083 (8)0.0026 (8)
C20.0501 (14)0.0552 (14)0.0690 (15)0.0191 (12)0.0049 (11)0.0064 (11)
C30.0898 (19)0.0674 (16)0.0493 (13)0.0514 (15)0.0007 (12)0.0056 (11)
C40.0518 (12)0.0460 (12)0.0365 (10)0.0264 (10)0.0029 (9)0.0014 (9)
C50.0577 (13)0.0484 (12)0.0364 (10)0.0303 (11)0.0001 (9)0.0009 (9)
C60.0535 (12)0.0444 (11)0.0311 (9)0.0297 (10)0.0052 (8)0.0026 (8)
C70.0452 (11)0.0423 (11)0.0400 (10)0.0223 (10)0.0024 (9)0.0009 (8)
C80.0480 (12)0.0543 (13)0.0379 (10)0.0317 (11)0.0011 (9)0.0045 (9)
C90.0486 (12)0.0431 (11)0.0413 (10)0.0291 (10)0.0088 (9)0.0080 (8)
C100.0491 (13)0.0422 (12)0.0598 (13)0.0203 (10)0.0055 (10)0.0006 (10)
C110.0566 (14)0.0538 (13)0.0468 (11)0.0310 (11)0.0109 (10)0.0005 (10)
Geometric parameters (Å, º) top
O1—C11.424 (2)C3—H3C0.9600
O1—H1O0.8200C4—C51.189 (3)
O2—N11.221 (3)C5—C61.438 (3)
O3—N11.219 (2)C6—C111.387 (3)
N1—C91.466 (3)C6—C71.390 (3)
C1—C41.471 (3)C7—C81.382 (3)
C1—C21.516 (3)C7—H70.9300
C1—C31.523 (3)C8—C91.371 (3)
C2—H2A0.9600C8—H80.9300
C2—H2B0.9600C9—C101.376 (3)
C2—H2C0.9600C10—C111.375 (3)
C3—H3A0.9600C10—H100.9300
C3—H3B0.9600C11—H110.9300
C1—O1—H1O109.5H3B—C3—H3C109.5
O3—N1—O2123.3 (2)C5—C4—C1175.7 (2)
O3—N1—C9118.5 (2)C4—C5—C6176.5 (2)
O2—N1—C9118.2 (2)C11—C6—C7119.25 (18)
O1—C1—C4106.76 (15)C11—C6—C5120.46 (18)
O1—C1—C2109.11 (16)C7—C6—C5120.27 (19)
C4—C1—C2110.61 (18)C8—C7—C6120.36 (19)
O1—C1—C3110.10 (17)C8—C7—H7119.8
C4—C1—C3108.98 (16)C6—C7—H7119.8
C2—C1—C3111.20 (18)C9—C8—C7118.75 (18)
C1—C2—H2A109.5C9—C8—H8120.6
C1—C2—H2B109.5C7—C8—H8120.6
H2A—C2—H2B109.5C8—C9—C10122.25 (18)
C1—C2—H2C109.5C8—C9—N1118.80 (18)
H2A—C2—H2C109.5C10—C9—N1118.95 (19)
H2B—C2—H2C109.5C11—C10—C9118.6 (2)
C1—C3—H3A109.5C11—C10—H10120.7
C1—C3—H3B109.5C9—C10—H10120.7
H3A—C3—H3B109.5C10—C11—C6120.79 (19)
C1—C3—H3C109.5C10—C11—H11119.6
H3A—C3—H3C109.5C6—C11—H11119.6
C11—C6—C7—C80.7 (3)O3—N1—C9—C109.2 (3)
C5—C6—C7—C8177.66 (17)O2—N1—C9—C10170.3 (2)
C6—C7—C8—C90.2 (3)C8—C9—C10—C111.2 (3)
C7—C8—C9—C101.2 (3)N1—C9—C10—C11178.21 (19)
C7—C8—C9—N1178.28 (17)C9—C10—C11—C60.3 (3)
O3—N1—C9—C8171.3 (2)C7—C6—C11—C100.7 (3)
O2—N1—C9—C89.2 (3)C5—C6—C11—C10177.72 (19)
Hydrogen-bond geometry (Å, º) top
D—H···AD—HH···AD···AD—H···A
O1—H1O···O1i0.821.872.682 (2)173
C10—H10···O3ii0.932.673.548 (3)157
C11—H11···O2iii0.932.683.467 (3)143
Symmetry codes: (i) xy+1/3, x1/3, z+5/3; (ii) x+y+2/3, x+4/3, z+1/3; (iii) y+4/3, xy+2/3, z+2/3.
Summary of short interatomic contacts (Å) in (I) top
The interatomic distances are calculated in Crystal Explorer 17 (Turner et al., 2017) whereby the X—H bond lengths are adjusted to their neutron values.
ContactDistanceSymmetry operation
O1···H3A2.711/3 + y, 2/3 - x + y, 5/3 - z
O2···H2B2.692/3 - y, 1/3 + x - y, -2/3 + z
O3···H2A2.691 - x, 1 - y, 1 - z
C1···H1O2.851/3 + y, 2/3 - x + y, 5/3 - z
C5···H3C2.791/3 + y, 2/3 - x + y, 2/3 - z
C7···H2C2.851/3 + y, 2/3 - x + y, 2/3 - z
C8···H2C2.801/3 + y, 2/3 - x + y, 2/3 - z
Percentage contributions of interatomic contacts to the Hirshfeld surface for (I) top
ContactPercentage contribution
H···H38.2
O···H/H···O32.1
C···H/H···C20.0
C···C4.2
N···O/O···N1.7
O···O1.6
C···N/N···C1.0
N···H/H···N0.8
C···O/O···C0.4
Summary of interaction energies (kJ mol-1) calculated for (I) top
ContactR (Å)EeleEpolEdisErepEtot
O1—H1O···O1i
H3A···O1i8.80-52.3-12.0-18.872.7-35.7
H1O···C1i
C10—H10···O3ii8.28-3.7-1.4-9.24.9-9.8
C11—H11···O2iii9.51-5.8-1.7-5.75.0-9.6
O3···H2Aiv
(C6–C11)···(C6–C11)iv4.25-9.4-1.8-47.128.9-34.4
H3C···C5v
H2C···C7v
H2C···C8v5.78-2.1-0.7-28.618.2-16.4
C2—H2C···(C6–C11)v
Symmetry codes: (i) 1/3 + x - y, 1/3 - x, 5/3 - z; (ii) 2/3 - x + y, 4/3 - x, 1/3 + z; (iii) 4/3 - y, 2/3 + x - y, 2/3 + z; (iv) 1 - x, 1 - y, 1 - z; (v) 1/3 + x - y, - 1/3 + x, 2/3 - z.
Geometric data (Å, °) for related 2-methyl-4-(aryl)but-3-yn-2-ol molecules top
CompoundZ'Cring—CacetyleneCacetylene—CacetyleneCacetylene—Cquaternarysupramolecular motifReference
(I)11.438 (3)1.189 (3)1.471 (3)hexamerThis work
(II)31.443 (5)1.211 (5)1.454 (5)chainSingelenberg & van Eijck (1987)
1.437 (6)1.192 (6)1.479 (6)
1.437 (5)1.189 (5)1.479 (5)
(III)21.441 (2)1.193 (2)1.490 (2)dimerClegg (2017)
1.435 (2)1.1895 (2)1.480 (2)
(IV)11.4377 (16)1.2000 (16)1.4791 (16)chainEissmann et al. (2010)
(V)31.4418 (18)1.1951 (19)1.4764 (19)hexamerHübscher et al. (2016)
1.444 (2)1.194 (2)1.4859 (19)
1.4402 (19)1.1904 (19)1.4723 (18)
 

Footnotes

Additional correspondence author, e-mail: julio@power.ufscar.br.

Acknowledgements

We thank Professor Regina H. A. Santos from IQSC-USP for the X-ray data collection.

Funding information

Funding for this research was provided by GlaxoSmithKline (GSK) and the Brazilian agencies: The National Council for Scientific and Technological Development are thanked for fellowships (CNPq: 308480/2016-3 to IC; 303207/2017-5 to JZS), São Paulo Research Foundation (FAPESP, grants 2013/06558-3 and 2014/50249-8) and Coordenação de Aperfeiçoamento de Pessoal de Nível Superior – Brasil (CAPES) – Finance Code 001. Sunway University Sdn Bhd is also thanked for funding (grant No. STR-RCTR-RCCM-001-2019).

References

First citationBleicher, L. S., Cosford, N. D. P., Herbaut, A., McCallum, J. S. & McDonald, I. A. (1998). J. Org. Chem. 63, 1109–1118.  CrossRef CAS Google Scholar
First citationBrandenburg, K. (2006). DIAMOND. Crystal Impact GbR, Bonn, Germany.  Google Scholar
First citationBruker (2009). APEX2 and SAINT. Bruker AXS Inc., Madison, Wisconsin, USA.  Google Scholar
First citationBurla, M. C., Caliandro, R., Carrozzini, B., Cascarano, G. L., Cuocci, C., Giacovazzo, C., Mallamo, M., Mazzone, A. & Polidori, G. (2015). J. Appl. Cryst. 48, 306–309.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationChemAxon (2010). Marvinsketch. https://www.chemaxon.com.  Google Scholar
First citationClegg, W. (2017). Private communication (refcode: HEFDAA). CCDC, Cambridge, England.  Google Scholar
First citationContreras-García, J., Johnson, E. R., Keinan, S., Chaudret, R., Piquemal, J.-P., Beratan, D. N. & Yang, W. (2011). J. Chem. Theory Comput. 7, 625–632.  Web of Science PubMed Google Scholar
First citationEissmann, F., Kafurke, U. & Weber, E. (2010). Acta Cryst. E66, o1866.  CSD CrossRef IUCr Journals Google Scholar
First citationErdélyi, M. & Gogoll, A. (2001). J. Org. Chem. 66, 4165–4169.  PubMed Google Scholar
First citationFarrugia, L. J. (2012). J. Appl. Cryst. 45, 849–854.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationHübscher, J., Rosin, R., Seichter, W. & Weber, E. (2016). Acta Cryst. E72, 1370–1373.  CSD CrossRef IUCr Journals Google Scholar
First citationHundertmark, T., Littke, A. F., Buchwald, S. L. & Fu, G. C. (2000). Org. Lett. 2, 1729–1731.  CrossRef PubMed CAS Google Scholar
First citationJohnson, E. R., Keinan, S., Mori-Sánchez, P., Contreras-García, J., Cohen, A. J. & Yang, W. (2010). J. Am. Chem. Soc. 132, 6498–6506.  Web of Science CrossRef CAS PubMed Google Scholar
First citationLi, X., Sun, S., Yang, F., Kang, J., Wu, Y. & Wu, Y. (2015). Org. Biomol. Chem. 13, 2432–2436.  CrossRef CAS PubMed Google Scholar
First citationMcKinnon, J. J., Jayatilaka, D. & Spackman, M. A. (2007). Chem. Commun. pp. 3814–3816.  Web of Science CrossRef Google Scholar
First citationSheldrick, G. M. (1996). SADABS. University of Göttingen, Germany.  Google Scholar
First citationSheldrick, G. M. (2015). Acta Cryst. C71, 3–8.  Web of Science CrossRef IUCr Journals Google Scholar
First citationSingelenberg, F. A. J. & van Eijck, B. P. (1987). Acta Cryst. C43, 693–695.  CSD CrossRef CAS Web of Science IUCr Journals 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 citationTan, X., Kong, L., Dai, H., Cheng, X., Liu, F. & Tschierske, C. (2013). Chem. Eur. J. 19, 16303–16313.  CrossRef CAS PubMed Google Scholar
First citationTurner, M. J., Mckinnon, J. J., Wolff, S. K., Grimwood, D. J., Spackman, P. R., Jayatilaka, D. & Spackman, M. A. (2017). Crystal Explorer 17. The University of Western Australia.  Google Scholar
First citationWestrip, S. P. (2010). J. Appl. Cryst. 43, 920–925.  Web of Science CrossRef CAS IUCr Journals 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