Next Article in Journal
Inorganic Amino-Nitro-Guanidinium Derivatives
Previous Article in Journal
Single-Component Charge-Transfer Crystals Based on Spin-Carrying TCNQ (7,7,8,8-Tetracyanoquinodimethane) Derivatives
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Non-Covalent Interactions in the Crystal Structure of Methyl 4-Hydroxy-3-Nitrobenzoate

1
Changsha Medical University, Changsha 410219, China
2
School of Chemistry & Biological Engineering, Changsha University of Science and Technology, Changsha 410114, China
3
Department of Chemistry, University of Otago, P. O. Box 56, Dunedin 9054, New Zealand
*
Author to whom correspondence should be addressed.
Crystals 2012, 2(2), 669-674; https://doi.org/10.3390/cryst2020669
Submission received: 18 April 2012 / Revised: 22 May 2012 / Accepted: 1 June 2012 / Published: 12 June 2012

Abstract

:
Methyl 4-hydroxy-3-nitrobenzoate, (I), C8H7NO5, crystallizes with two unique molecules, A and B, in the asymmetric unit of the triclinic unit cell. The space group was assigned as P-1, with lattice parameters a = 0.72831(15), b = 1.0522(2), c = 1.1410(2) nm, α = 83.38(3), β = 80.83(3), γ = 82.02(3)°, Z = 4, V = 0.8510(3) nm3, Mr = 197.15, Dc = 1.539 g/m3, µ= 0.131 mm−1, F(000) = 408, R = 0.1002 and wR = 0.2519. In the crystal structure, 12 hydrogen bonding and two π-stacking interactions link the molecules into infinite stacked sheets parallel to (101).

1. Introduction

Weak noncovalent interactions play a significant role in biological or biomimetic systems as well as in artificial supramolecular structures. They can stabilize the three-dimensional structure of large molecules, such as proteins and nucleic acids, and are involved in many biological processes [1]. Also, the noncovalent bond is the dominant contact between supramolecules in supramolecular chemistry [2]. These forces commonly include both classical and non-classical hydrogen bonding, π-π and C—H...π interactions together with hydrophobic forces, van der Waals interactions, and electrostatic effects [2]. The study of these non-covalent interactions is crucial to understanding the structures of organic crystals, drug binding, and many biological processes. Experimentally, these interactions are often seen in complex environments, where it can be difficult to pick out only the interaction of interest. Theoretically, they feature shallow potential energy surfaces and require very accurate quantum-mechanical modeling for reliable predictive results. In the structure of methyl 4-hydroxy-3-nitrobenzoate, C8H7NO5 (I), determined by single crystal X-ray analysis, we examine the variety of non-covalent interactions that lead to the overall stability of the crystal structure.

2. Results and Discussion

Methyl 4-hydroxy-3-nitrobenzoate (I) crystallizes with two unique molecules in the asymmetric unit of the triclinic unit cell (Figure 1). The two molecules are closely similar and overlay with an rms deviation of 0.0193 Å and a maximum deviation of 0.0398 Å [3]. Both the N1 O1 O2 nitro and C7 O4 O5 C8 methyl ester substituents lie close to the planes of the benzene rings in both molecules, with dihedral angles between the planes of 6.1 (6) ° and 10.5 (3) ° in molecule A and 6.2 (7) ° and 10.4 (2) ° in molecule B. Intramolecular O3A—H3A…O2A and O3B—H3B…O2B hydrogen bonds, Table 1, generate planar S(6) rings [4]. Rms deviations from these ring planes are 0.035Å for molecule A and 0.023Å for molecule B suggesting that these contacts contribute to the overall planarity of both molecules. A search of the Cambridge Database [5] reveals two isomeric forms of (I), namely methyl 2-hydroxy-3-nitrobenzoate (II) [6] and methyl 2-hydroxy-5-nitrobenzoate (III) [7] together with the closely related methyl 2-hydroxy-3,5-dinitrobenzoate (IV) [8]. In all three similar structures, intramolecular O—H...O hydrogen bonds involve the OH substituent and either a nitro or carbonyl oxygen and ensure that the molecules adopt reasonably planar conformations. Interestingly, for (II) and (IV), where a choice of O acceptor atom is possible due to the relative placements of the nitro and methyl ester substituents, the carbonyl O atom of the methyl ester is the preferred acceptor. In addition, the structures of a number of derivatives in which the phenol H atom has been replaced to form alkyl [9] and aryl [10,11] ether derivatives have also been reported.
Figure 1. The asymmetric unit of (I) with displacement ellipsoids for the non-hydrogen atoms drawn at the 50% probability level, and intramolecular hydrogen bonds are drawn as dashed lines in molecules A and B.
Figure 1. The asymmetric unit of (I) with displacement ellipsoids for the non-hydrogen atoms drawn at the 50% probability level, and intramolecular hydrogen bonds are drawn as dashed lines in molecules A and B.
Crystals 02 00669 g001
Table 1. Hydrogen-bond geometries for compound I.
Table 1. Hydrogen-bond geometries for compound I.
BondD–HH…AD…AD–H…A
O3A—H3A…O2A0.828(11)1.94(4)2.585 (5)134 (5)
O3B—-H3B…O2B0.831(11)1.91 (4)2.584 (5)137 (5)
C8A—H8AB…O3A i0.962.623.372 (7)136
C8A—H8AA…O3A ii0.962.613.441 (7)145
C4A—H4A…O4B iii0.932.413.136 (7)135
O3A—H3A…O1B iv0.828 (11)2.30 (5)2.896 (5)129 (5)
C1A—H1A…O2B v0.932.563.483 (6)174
C4B—H4B…O4A vi0.932.493.327 (6)150
C1B—H1B…O2A iv0.932.593.509 (6)168
O3B—H3B…O1A v0.831 (11)2.28 (4)2.863 (5)128 (5)
C8B—H8BB…O3B vii0.962.503.282 (6)139
C8B—H8BA…O3B viii0.962.703.475 (7)139
Symmetry codes: i x, y+1, z; iix, −y + 1, −z + 1; iiix, −y, −z + 2; ivx + 1, −y, −z + 1; vx + 1, −y + 1, −z + 1; vix, −y + 1, −z + 2; viix, y − 1, z; viiix + 1, −y, −z + 2.
In the crystal structure two sets of centrosymmetric dimers form through O3A—H3A…O1B hydrogen bonds together with C1B—H1B…O2A interactions and O3B—H3B…O1A and C1A—H1A…O2B contacts (Figure 2). The formation of each dimer generates R22 (11) rings [4]. Additional weak C—H…O contacts link these dimers into sheets in the bc plane. π-Stacking interactions are also found between the aromatic rings of adjacent molecules, linking them in a head to tail fashion (Figure 3). Centroid to centroid distances are 3.713 (3) for the A and 3.632 (3) for the B molecules. These interactions are augmented by additional weak inversion related C8A—-H8AA…O3A and C8B—H8BA…O3B contacts forming AA and BB dimers and generating R22(18) rings. The two sets of dimers combine to form sheets parallel to the (101) plane (Figure 4). Similar sheet formation is also common to the crystal structures of the isomeric forms (II) and (III) [6,7]; significant C—H...O hydrogen bonds are found in both structures forming layers that are further connected by weak π...π contacts.
Figure 2. Centrosymmetric dimers forming sheets in the bc plane with hydrogen bonds drawn as dashed lines.
Figure 2. Centrosymmetric dimers forming sheets in the bc plane with hydrogen bonds drawn as dashed lines.
Crystals 02 00669 g002
Figure 3. Centrosymmetric dimers formed by π…π contacts (dotted lines) and weak C—H…O interactions dashed lines. The red spheres represent centroids of the benzene rings in the A molecules and the blue sphere those in the B molecules.
Figure 3. Centrosymmetric dimers formed by π…π contacts (dotted lines) and weak C—H…O interactions dashed lines. The red spheres represent centroids of the benzene rings in the A molecules and the blue sphere those in the B molecules.
Crystals 02 00669 g003
Figure 4. Crystal packing for (I) viewed along the b axis with hydrogen bonds drawn as dashed lines.
Figure 4. Crystal packing for (I) viewed along the b axis with hydrogen bonds drawn as dashed lines.
Crystals 02 00669 g004

3. Experimental Section

Synthesis of methyl 4-hydroxy-3-nitrobenzoate (I). Cerium (IV) ammonium nitrate (CAN) (10.96 g, 20 mmol) was added to a mixture of methyl 4-hydroxybenzoate (1.5 g, 10 mmol) and NaHCO3 (3.5 g) in anhydrous MeCN (60 mL) at room temperature while stirring. The reaction mixture was then stirred at room temperature for a further 30 min. The resulting mixture was filtered, washed with water, and extracted with ethyl acetate (3 × 25 mL). The combined extracts were dried over anhydrous Na2SO4, and the solvent removed to furnish the title compound (70%). m.p. 74–76 °C. 1H NMR (DMSO-d6, 400 MHz, ppm) δ 8.39 (d, J = 1.7 Hz, 1 H), 8.06 (dd, J = 8.7 Hz, 1.7 Hz, 1 H), 7.23 (dd, J = 8.7 Hz, 1.7 Hz, 1 H), 3.85 (s, 3 H). 13C NMR (DMSO-d6,100 MHz, ppm) 165.4, 156.7, 137.5, 136.0, 127.6, 121.2, 120.2, 53.0. LC/MS: m/z 196 ([M − H]). Anal. calcd (%). for C8H7NO5: C 48.74, H 3.58, N 7.10. Found (%): C 48.66, H 3.51, N 7.03. Colorless plate-like crystals of (I) were grown from a solution of petroleum ether and ethyl acetate (1:1, v/v) by slow natural evaporation at room temperature.
X-ray Data Collection and Structure Solution. A single crystal of (I) with dimensions 0.20 mm × 0.18 mm × 0.04 mm was selected for data collection performed on a Rigaku Saturn CCD area detector (λ = 0.71073 Å) [12] using the ω-2θ scan mode. A total of 5670 reflections were collected in the range 1.8 ≤ θ ≤ 26.1°. Of these 2959 were independent (Rint = 0.129), and 1408 were considered to be observed with I > 2σ(I) and used in the succeeding refinement. The structure was solved by direct methods and refined on F2 by full matrix least-squares using the SHELXTL program [13]. The H3A and H3B atoms of the hydroxy groups were located in a difference Fourier map and their coordinates refined with Uiso = 1.5Ueq (O). All other H-atoms were refined using a riding model with d(C—H) = 0.93Å, Uiso=1.2Ueq (C) for aromatic and 0.96Å, Uiso = 1.5Ueq (C) for CH3 H atoms. Goodness-of-fit on F2 is 0.951. R = 0.1003 and wR2 = 0.2519, with (w = 1/[σ2(Fo2) + (0.0965P)2], where P = (Fo2 + 2Fc2)/3). (Δρ)max = 0.43, (Δρ)min = –0.45 e/Å3. The crystal was very weakly diffracting with less than half of the measured reflections with I > 2σ(I) and a high R(int). In addition, anisotropic refinement led to elongations of displacement ellipsoids for some atoms and restraints were used in order to minimise these effects. The final residuals are therefore somewhat higher than normal as a result, but this does not markedly affect the integrity of the overall structure.

4. Conclusions

The molecular and crystal structure of methyl 4-hydroxy-3-nitrobenzoate (I) is reported. The structure is stabilised by an extensive set of twelve hydrogen bonding and two π-stacking interactions.

Acknowledgments

Financial support from the Scientific Research Fund of Hunan Provincial Education Department (10B012 & 10W017) is gratefully acknowledged.

References

  1. Lodish, H.; Berk, A.; Kaiser, C.A.; Krieger, M.; Scott, M.P.; Bretscher, A.; Ploegh, H.; Matsudaira, P. Molecular Cell Biology, 6th ed; W.H. Freeman Publishers: New York, NY, USA, 2007; pp. 29–56. [Google Scholar]
  2. Steed, J.W.; Atwood, J.L. Supramolecular Chemistry, 2nd ed; John Wiley & Sons Ltd.: Chichester, UK, 2000; pp. 27–36. [Google Scholar]
  3. Macrae, C.F.; Bruno, I.J.; Chisholm, J.A.; Edgington, P.R.; McCabe, P.; Pidcock, E.; Rodriguez-Monge, L.; Taylor, R.; van de Streek, J.; Wood, P.A. Mercury CSD 2.0-New features for the visualization and investigation of crystal structures. J. Appl. Cryst. 2008, 41, 466–470. [Google Scholar] [CrossRef]
  4. Bernstein, J.; Davis, R.E.; Shimoni, L.; Chang, N.-L. Patterns in hydrogen bonding: Functionality and graph set analysis in crystals. Angew. Chem. Int. Ed. Engl. 1995, 34, 1555–1573. [Google Scholar] [CrossRef]
  5. Allen, F.H. The Cambridge structural database: A quarter of a million crystal structures and rising. ActaCryst. 2002, B58, 380–388. [Google Scholar]
  6. Liu, Y.-Z.; Li, Y.-X.; Zhang, L.; Li, X. Methyl 2-hydroxy-3-nitrobenzoate. Acta Cryst. 2009, E65. [Google Scholar]
  7. Jin, L.-F.; Xiao, F.-P. Methyl 5-nitrosalicylate. Acta Cryst. 2005, E61, o1107–o1108. [Google Scholar]
  8. Jin, L.-F.; Xiao, F.-P.; Wang, C.-G. Methyl 3,5-dinitrosalicylate. Acta Cryst. 2004, E60, o1593–o1594. [Google Scholar]
  9. Zhao, S.-Y.; Mou, S.-W.; Zhao, S.-H.; Qin, W.-M. Methyl 3-nitro-4-propoxybenzoate. Acta Cryst. 2007, E63. [Google Scholar]
  10. Gopal, R.; Chandler, W.D.; Robertson, B.E. Conformations of bridged diphenyls. XV. Crystal structure of (4-carbomethoxy-2-nitrophenoxy)benzene and the effect of intramolecular interactions in nitro substituents. Can. J. Chem. 1980, 58, 658–663. [Google Scholar] [CrossRef]
  11. Van der Heijden, S.P.N.; Griffith, E.A.H.; Chandler, W.D.; Robertson, B.E. Conformations of bridged diphenyls. VII. Crystal structure of 2-(4′-Carbomethoxy-2′-nitrophenoxy)-1,3,5-trimethylbenzene. Can. J. Chem. 1975, 53, 2084–2092. [Google Scholar] [CrossRef]
  12. Rigaku/MSC, CrystalClear; Rigaku/MSC Inc.: The Woodlands, TX, USA, 2005.
  13. Sheldrick, G.M. A short history of SHELX. ActaCryst. 2008, A64, 112–122. [Google Scholar]

Share and Cite

MDPI and ACS Style

Fu, X.-L.; Li, J.-S.; Simpson, J. Non-Covalent Interactions in the Crystal Structure of Methyl 4-Hydroxy-3-Nitrobenzoate. Crystals 2012, 2, 669-674. https://doi.org/10.3390/cryst2020669

AMA Style

Fu X-L, Li J-S, Simpson J. Non-Covalent Interactions in the Crystal Structure of Methyl 4-Hydroxy-3-Nitrobenzoate. Crystals. 2012; 2(2):669-674. https://doi.org/10.3390/cryst2020669

Chicago/Turabian Style

Fu, Xin-Ling, Jiang-Sheng Li, and Jim Simpson. 2012. "Non-Covalent Interactions in the Crystal Structure of Methyl 4-Hydroxy-3-Nitrobenzoate" Crystals 2, no. 2: 669-674. https://doi.org/10.3390/cryst2020669

Article Metrics

Back to TopTop