3,5,7-Trimethoxyphenanthrene-1,4-dione: a new biologically relevant natural phenanthrenequinone derivative from Dioscorea prazeri and studies on its single X-ray crystallographic behavior, molecular docking and other physico-chemical properties

Goutam Brahmachari*a, Saktipada Dasb, Maya Biswas (Sinha)c, Abhishek Kumard, Ambrish Kumar Srivastavad and Neeraj Misrad
aLaboratory of Natural Products & Organic Synthesis, Department of Chemistry, Visva-Bharati, (Central University), Santiniketan-731 235, West Bengal, India. E-mail: brahmg2001@yahoo.co.in; brahmg2001@gmail.com; Fax: +91-3463-526; Tel: +91-3463-526
bDepartment of Chemistry, University of Kalyani, Kalyani-741 235, West Bengal, India
cDepartment of Chemistry, Krishnagar Women's College, Nadia-741101, West Bengal, India
dDepartment of Physics, University of Lucknow, Lucknow-226007, Uttar Pradesh, India

Received 15th October 2015 , Accepted 5th January 2016

First published on 8th January 2016


Abstract

A naturally occurring new phenanthrenequinone derivative has been isolated from the yams of Dioscorea prazeri Prain and Burkil (Dioscoreaceae) and identified as 3,5,7-trimethoxyphenanthrene-1,4-dione (1) on the basis of its detailed spectral and single crystal X-ray analyses. The compound crystallizes in the triclinic space group P[1 with combining macron] with the following unit-cell parameters: a = 7.7045(3), b = 8.5088(4), c = 16.3254(7) Å, α = 98.908(2)°, β = 96.316(2)°, γ = 103.939(2)° and Z = 2. The crystal structure was solved by direct methods using single-crystal X-ray diffraction data collected at room temperature and refined by full-matrix least-squares procedures to a final R-value of 0.0559 for 2832 observed reflections. Exhaustive theoretical studies on the molecular structure, vibrational spectra, HOMO, LUMO, MESP surfaces, reactivity descriptor and molecular docking of this plant-derived hitherto unknown natural molecule have also been performed. The equilibrium geometry of the title compound has been obtained and analyzed using the DFT-B3LYP/6-31+G(d,p) method. A comparison between the calculated vibrational spectral data with the experimental observations for the phenanthrenequinone molecule has been performed. The reactivity of this molecule is explained using various local as well as global molecular descriptors and reactivity surfaces have also been analyzed. The molecular docking study of the title molecule for predicting its possible anticancer property has also been investigated.


1. Introduction

Dioscorea prazeri Prain and Burkil (family: Dioscoreaceae), locally known as Kukur torul, grows abundantly in the wetter parts of the Eastern Himalayas in North Bengal, Sikkim, Bhutan and in the Naga hills up to 5500 ft height.1 It is an important medicinal plant well-known as a source of the potentially bioactive compound, diosgenin.2 Previous phytochemical investigations on this plant yielded several steroidal sapogenins and saponins such as prazerigenins A–D,3 smilagenone and epismilagenone,4 diosgenin-3-O-α-L-rhamnopyranosyl(1 → 6)-β-D-glucopyranoside, diosgenin-3-O-α-L-rhamnopyranosyl(1 → 6)-β-D-glucopyranosyl(1 → 6)-β-D-glucopyranoside and prazerigenin A-3-O-α-L-rhamnopyranosyl(1 → 6)-β-D-glucopyranosyl(1 → 6)-β-D-glucopyranoside5 along with a few phenanthrene derivatives such as 5,6-dihydroxy-2,3,4-trimethoxy-9,10-dihydrophenanthrene (prazerol),6 5,6-dihydroxy-1,3,4-trimethoxy-9,10-dihydrophenanthrene and 5,6-dihydroxy-2,4-dimethoxy-9,10-dihydrophenanthrene.7 In continuation on our studies on natural products,8 we isolated a new natural phenanthrenquinone derivative identified as 3,5,7-trimethoxyphenanthrene-1,4-dione (1) from the yams of D. prazeri. In this communication, we report the isolation, structural elucidation based on detailed spectral and X-ray analyses and combined experimental and theoretical studies on the molecular structure, vibrational spectra, HOMO, LUMO, MESP surfaces, reactivity descriptor and molecular docking of this plant-derived new natural molecule.

2. Experimental

2.1 Chemicals and instrumentation

All the chemicals used in the synthesis were of analytical grade and used without purification. All the solvents were dried before use by the literature methods.9

Melting point was recorded on a Chemiline CL-726 melting point apparatus and is uncorrected. The infrared spectra were recorded on FT-IR-8400S using KBr disc. 1H and 13C NMR spectra were obtained at 400 MHz and 100 MHz, respectively, on a Bruker DRX spectrometer with CDCl3 as the solvent. TMS was used as internal standard in recording NMR spectra. Mass spectra (TOF-MS) were measured on a QTOF Micro mass spectrometer. Elemental analyses were performed on an Elementar Vario EL III Carlo Erba 1108 microanalyzer instrument. Chromatography was carried on silica gel columns (Merck 60–120 mesh) and TLC was performed on silica gel 60 F254 (Merck) plates. X-ray crystallographic data for the compound were collected on Bruker SMART X2S Single Crystal X-ray diffractometer equipped with graphite monochromated MoKα radiation (λ = 0.71073 Å).

2.2 Extraction and isolation of phenanthrenequinone 1

Dried and powdered yams (∼2.2 kg) of Dioscrorea prazeri were extracted in a Soxhlet apparatus with petroleum ether (60–80 °C) for 8 h. Evaporation of the solvent gave a red gummy residue (∼40 g) which was dissolved in ether and washed with 4% aq. KOH solution to separate the acidic components present in the extract. The organic layer on usual work up afforded a brown-red gummy residue (∼12 g) which was subjected to chromatographic resolution over silica gel (mesh 100–200). Petrol–benzene (1[thin space (1/6-em)]:[thin space (1/6-em)]4) eluents on concentration and standing under refrigeration yielded an orange-red solid (60 mg, mp 152–154 °C) that crystallized from a mixture of chloroform and methanol to furnish needle-shaped orange-red crystals of the phenanthrenequinone derivative 1.

3,5,7-Trimethoxyphenanthrene-1,4-dione (1). Yield: 48 mg (0.0024%), orange-red needles, mp 166–68 °C; IR, 1H NMR (400 MHz, CDCl3), 13C NMR (100 MHz, CDCl3), DEPT-135, 1H–1H COSY, 1H–13C HMQC, and 1H–13C HMBC data are described in the text (also in Table 1); HR-TOF-MS: m/z 321.0732 (C17H14O5Na, [M + Na]+ requires 321.0739). Anal. calcd for C17H14O5: C, 68.45; H, 4.73; found C, 68.43; H, 4.78. For single crystal X-ray analyses, 10 mg of compound was dissolved in 5 mL chloroform and left for several days at ambient temperature to yield orange-red block shaped crystals (Fig. 1).

Table 1 1D and 2D-NMR spectral behavior of 3,5,7-trimethoxyphenanthrene-1,4-dione (1)
Carbon 1H (ppm/δ) 13C (ppm/δ) DEPT-135 1H–1H COSY-45 1H–13C HMQC 1H–13C HMBC
1 181.74 C
2 6.00 (1H, s) 106.18 CH δ 6.00 (H-2) vs. δ 106.18 (C-2) δ 6.00 (H-2) vs. δ 162.83 (C-3), 130.73 (C-10a)
3 162.83 C
4 184.59 C
4a 138.95 C
4b 116.84 C
5 158.05 C
6 6.69 (1H, d, J = 2.4 Hz) 101.92 CH δ 6.69 (H-6) vs. δ 101.92 (C-6) δ 6.69 (H-6) vs. δ 116.84 (C-4b), 160.76 (C-7)
7 160.76 C
8 6.76 (1H, d, J = 2.4 Hz) 99.06 CH δ 6.76 (H-8) vs. δ 99.06 (C-8) δ 6.76 (H-8) vs. δ 101.92 (C-6), 116.84 (C-4b), 132.27 (C-9)
8a 132.70 C
9 7.87 (1H, d, J = 8.8 Hz) 132.27 CH H-9 (δ 7.87) vs. H-10 (δ 8.04) δ 7.87 (H-9) vs. δ 132.27 (C-9) δ 7.87 (H-9) vs. δ 138.94 (C-4a), 116.84 (C-4b), 99.06 (C-8), 130.73 (C-10a)
10 8.04 (1H, d, J = 8.4 Hz) 122.53 CH H-10 (δ 8.04) vs. H-9 (δ 7.87) δ 8.04 (H-9) vs. δ 122.53 (C-10) δ 8.04 (H-10) vs. δ 138.94 (C-4a), 132.27 (C-9)
10a 130.73 C  
C3–OCH3 3.94 (s) 56.48 CH3 δ 3.94 (OCH3) vs. δ 56.48 (OCH3) δ 3.94 (C3–OCH3) vs. δ 162.83 (C-3)
C5–OCH3 3.94 (s) 55.54 CH3 δ 3.94 (OCH3) vs. δ 55.54 (OCH3) δ 3.94 (C5–OCH3) vs. δ 158.05 (C-5)
C7–OCH3 3.94 (s) 55.87 CH3 δ 3.94 (OCH3) vs. δ 55.87 (OCH3) δ 3.94 (C3–OCH3) vs. δ 160.78 (C-7)



image file: c5ra21490d-f1.tif
Fig. 1 Chemical structure of 3,5,7-trimethoxyphenanthrene-1,4-dione (1).

2.3 Single crystal X-ray diffraction studies: crystal structure determination and refinement

X-ray intensity data of 26[thin space (1/6-em)]240 reflections (of which 3644 unique) were collected on Bruker SMART X2S Single Crystal X-ray diffractometer equipped with graphite monochromated MoKα radiation (λ = 0.71073 Å). The crystal used for data collection was of dimensions 0.30 × 0.20 × 0.20 mm. The intensities were measured by ω scan mode for θ ranges 1.28 to 26.00°. 2832 reflections were treated as observed (I > 2σ(I)). Data were corrected for Lorentz, polarisation and absorption factors. The structure was solved by direct methods using SHELXS97.10 All non-hydrogen atoms of the molecule were located in the best E-map. Full-matrix least-squares refinement was carried out using SHELXL97.10 The final refinement cycles converged to an R = 0.0559 and wR(F2) = 0.1599 for the observed data. Residual electron densities ranged from −0.370 < Δρ < 0.392 e Å−3. Atomic scattering factors were taken from International Tables for X-ray Crystallography.11 The crystallographic data are summarized in Table 2, and the respective bond lengths and angles are shown in Table 3. ORTEP diagram and crystal packing for compound 1 are shown in Fig. 3 and 4. CCDC-1049913 contains the supplementary crystallographic data for this compound.12
Table 2 Crystal data and experimental details for 3,5,7-trimethoxyphenanthrene-1,4-dione (1)
CCDC number 1049913
Empirical formula C17H14O5·CHCl3
Formula weight 417.65
Radiation, wavelength Mo Kα, 0.71073 Å
Unit cell dimensions a = 7.0645(3), b = 8.5088(4), c = 16.3254(7) Å, α = 98.908(2)°, β = 96.316(2)°, γ = 103.939(2)°
Crystal size 0.3 × 0.2 × 0.2 mm
Crystal shape (color) Block shaped (orange-red)
Crystal system Triclinic
Space group P[1 with combining macron]
Unit cell volume 929.77(7)
No. of molecules per unit cell, Z 2
Temperature 293(2) K
Absorption coefficient 0.519 mm−1
F(000) 428
Scan mode ω scan
θ range for entire data collection 1.48 < θ < 26.00
Range of indices h = −8 to 8, k = −10 to 10, l = −20 to 20
Reflections collected/unique 26[thin space (1/6-em)]240/3644
Reflections observed (I > 2σ(I)) 2832
Rint 0.0420
Rsigma 0.0269
Structure determination Direct methods
Refinement Full-matrix least-squares on F2
No. of parameters refined 266
Final R 0.0559
wR(F2) 0.1599
Goodness-of-fit 1.086
Final residual electron density −0.370 < Δρ < 0.392 e Å−3
Measurement Bruker SMART X2S Single X-ray crystal diffractometer
Software for structure solution SHELXS97 [ref. 10]
Software for refinement SHELXL97 [ref. 10]
Software for molecular plotting WinGX,19 PLATON20
Software for geometrical calculation PLATON,20 PARST21


Table 3 Selected bond lengths and bond angles for 3,5,7-trimethoxyphenanthrene-1,4-dione (1)
Bond lengths
O3 C4 1.210(3) O2 C3 1.332(3)
O2 C12 1.437(3) O4 C5 1.352(3)
O4 C13 1.420(3) C4A C11 1.385(3)
C4A C4B 1.431(3) C4A C4 1.491(3)
C4B C8A 1.423(3) C4B C5 1.432(3)
C5 C6 1.365(3) C11 C10 1.408(3)
C11 C1 1.492(3) O5 C7 1.360(3)
O5 C14 1.423(4) C6 C7 1.405(4)
C4 C3 1.503(3) C1 O1 1.228(3)
C1 C2 1.447(4) C3 C2 1.337(3)
C7 C8 1.363(4) C8A C8 1.408(3)
C8A C9 1.414(3) C10 C9 1.356(4)
C20 Cl1 1.706(5) C20 Cl6 1.708(7)
C20 Cl3 1.722(7) C20 Cl5 1.725(6)
C20 Cl2 1.727(7) C20 Cl4 1.744(6)
[thin space (1/6-em)]
Bond angles
C3 O2 C12 117.00(18) C5 O4 C13 118.07(19)
C11 C4A C4B 119.9(2) C11 C4A C4 116.08(19)
C4B C4A C4 123.33(19) C8A C4B C4A 118.5(2)
C8A C4B C5 116.8(2) C4A C4B C5 124.6(2)
O4 C5 C6 123.2(2) O4 C5 C4B 116.24(19)
C6 C5 C4B 120.4(2) C4A C11 C10 120.2(2)
C4A C11 C1 121.0(2) C10 C11 C1 118.8(2)
C7 O5 C14 118.0(2) C5 C6 C7 121.0(2)
O3 C4 C4A 123.9(2) O3 C4 C3 119.3(2)
C4A C4 C3 116.42(19) O1 C1 C2 120.9(2)
O1 C1 C11 120.0(2) C2 C1 C11 119.0(2)
O2 C3 C2 127.2(2) O2 C3 C4 111.61(18)
C2 C3 C4 120.9(2) O5 C7 C8 125.7(2)
O5 C7 C6 113.8(2) C8 C7 C6 120.5(2)
C8 C8A C9 120.0(2) C8 C8A C4B 121.0(2)
C9 C8A C4B 1 19.0(2) C7 C8 C8A 119.6(2)
C3 C2 C1 119.9(2) C9 C10 C11 120.4(2)
C10 C9 C8A 121.4(2) Cl1 C20 Cl6 117.8(5)
Cl1 C20 Cl3 107.9(5) Cl1 C20 Cl5 100.4(4)
Cl6 C20 Cl5 107.0(6) Cl3 C20 Cl5 124.6(5)
Cl1 C20 Cl2 104.9(4) Cl6 C20 Cl2 92.5(6)
Cl3 C20 Cl2 110.0(6) Cl6 C20 Cl4 114.7(5)
Cl3 C20 Cl4 101.7(5) Cl5 C20 Cl4 113.1(5)
Cl2 C20 Cl4 118.1(5)    


2.4 Computational methods

All quantum chemical calculations of the title compound 1 are carried out with Gaussian 09 suite of program13 using the B3LYP/6-311++G(d,p) levels of theory to predict the molecular structure and vibrational wave numbers. The molecular geometry (Fig. 5) is fully optimized by Becke's three parameter exchange term combined with the gradient-corrected correlation functional of Lee, Yang and Parr.14 The B3LYP method is very popular for studying a variety of systems of biomolecular interests.15 The optimized structural parameters are used in the vibrational frequency calculations at DFT (B3LYP) levels. At the optimized geometry of the compound no imaginary frequency modes are obtained, therefore there is a true minimum on the potential energy surface.

3. Results and discussion

3.1 Structure elucidation of phenanthrenequinone derivative 1 based on spectral studies

Compound 1 was obtained as an orange-red solid (mp 166–168 °C) and having molecular formula C17H14O5 as deduced from its elemental analyses as well as from HR-TOF-MS ([M + Na]+, 321.0732). The UV absorption maxima of 1 at 223, 239 and 305 nm indicated the presence of a typical phenanthrene skeleton within the molecule.16 Seventeen carbon signals, including three methoxy, five methine and nine quaternary carbons, were observed in the 13C-NMR spectrum of 1; the DEPT-135 spectrum was also in complete agreement with this inference. Among the nine quaternary carbons, two were identified as carbonyl carbons on the basis of chemical shifts at δC 184.59 and δ 181.74 ppm. Therefore, the compound 1 was postulated to be a phenanthrenequinone. The IR spectrum of 1 showed characteristic absorption bands for the presence of α,β-unsaturated carbonyls attached with aromatic nucleus (3076, 1670, 1648, 1620, 1553, 1475, 1364 cm−1) and methoxy groups (2945, 2845, 1221, 1032 cm−1) into the molecule.

The 1H-NMR spectrum of 1 offered a clear indication about the substitution pattern of three fused rings (A, B and C) of the 1,4-phnanthrenequinone skeleton. The PMR spectrum displayed signals at (i) δ 8.04 (1H, d, J = 8.4 Hz) and δ 7.87 (1H, d, J = 8.8 Hz) ppm attributed to a pair of ortho-coupled C-ring adjacent aromatic protons attached to C-10 and C-9, respectively, (ii) δ 6.76 (1H, d, J = 2.4 Hz) and δ 6.69 (1H, d, J = 2.4 Hz) ppm were due to a pair of meta-coupled A-ring aromatic protons linked, respectively, to C-8 and C-6, (iii) δ 6.00 (1H, s) ppm for the only one olefinic proton attached with B-ring at C-2, and (iv) δ 3.94 (9H, s) ppm for three methoxyl functions, two of which are linked at C-5 and C-7 of ring A and the another at C-3 of ring B. These three methoxy carbons showed distinct resonances at δC 56.48, 55.87 and 55.54 ppm in the 13C-NMR spectrum. Thus, compound 1 may be postulated as 3,5,7-trimethoxy-1,4-phenanthrenequinone; the complete 13C-NMR spectral data and related DEPT-135 signals are in excellent agreement with those reported for compounds having similar skeleton.16,17

For further and thorough analysis, we performed its detailed 2D-NMR (1H–1H-COSY-45, HMQC and HMBC) studies. All respective homo- and heteronuclear interactions are shown in Fig. 2 and the results are summarized in Table 1. As expected, 1H–1H-COSY-45 spectrum of 1 showed only the interactions between H-9 and H-10 and vice versa. The HMQC spectrum of 1 also demonstrated the expected 1H–13C correlations between carbon atoms and the protons directly attached to them. Thus, H-2 (δ 6.00) correlates with C-2 (δ 106.18), H-6 (δ 6.69) with C-6 (δ 101.92), H-8 (δ 6.76) with C-8 (δ 99.06), H-9 (δ 7.87) with C-9 (δ 132.27), H-10 (δ 8.04) with C-10 (δ 122.53), and the methoxy protons (δ 3.94) with the three methoxy carbons at δ 56.48, 55.87 and 55.54 ppm. The results from the heteronuclear multiple bond correlation (HMBC) spectral studies unambiguously confirmed the structural pattern as proposed for the molecule 1. In the HMBC spectrum, C3–OCH3 protons showed interactions with C-3 at δ 162.83, while C5–OCH3 and C7–OCH3 protons exhibited such interactions, respectively with C-5 at δ 158.05 and C-7 at δ 160.78. H-2 was found to interact with C-3 at δ 162.83 and C-10a at δ 130.73. All other protons (H-6, H-8, H-9 and H-10) exhibited expected HMBC interactions as depicted in Fig. 2. The new compound 1 was, therefore, elucidated as 3,5,7-trimethoxyphenanthrene-1,4-dione. To our delight, we have been able to develop a unit crystal of the compound and performed its X-ray crystallographic studies that unequivocally established the structure of 1 as deduced based on spectral studies.


image file: c5ra21490d-f2.tif
Fig. 2 1H–1H COSY, 1H–13C HMQC, and 1H–13C HMBC interactions for 1 (COSY, correlation spectroscopy; HMQC, heteronuclear multiple quantum coherence; HMBC, heteronuclear multiple bond correlation).

3.2 X-ray analysis of phenanthrenequinone derivative 1

The crystal structure of the phenanthrenequinone 1 consists of three fused rings. An ORTEP view of the compound with atomic labeling is shown in Fig. 3.18 The geometry of the molecule was calculated using the WinGX,19 PARST20 and PLATON21 software. The overall molecular geometry of the compound, including bond distances has a normal range.22 The bond lengths C1–O1 and C4–O3 are 1.228(3) and 1.210(3), respectively, which are in conformity that these are double bonds. The C–O distances in methoxy groups attached to the rings are O2–C3 = 1.332(3), C5–O4 = 1.352(3) and O5–C7 = 1.360(3), respectively. In addition, all the three chlorine atoms in the chloroform solvent molecule are found disordered over two set of sites with occupancy ratio 0.525[thin space (1/6-em)]:[thin space (1/6-em)]0.475.
image file: c5ra21490d-f3.tif
Fig. 3 ORTEP view of the molecule with displacement ellipsoids drawn at 40%. H atoms are shown as small spheres of arbitrary radii.

Packing view of the molecule in the unit cell viewed down the a-axis is shown in Fig. 4a. In the crystal, C–H⋯O hydrogen bonds link the molecules into a two dimensional network. The intermolecular hydrogen bonds are responsible for the formation of hydrogen bonded network, thus, providing more stability to the molecules in the unit cell. A pair of intermolecular C12–H112⋯O1 hydrogen bonds link the molecules into inversion dimers generating R22(14) motif23 (Fig. 4b) linking the successive molecules. The crystal data and structure refinement details and selected bond lengths and angles are given in Tables 2 and 3, respectively.


image file: c5ra21490d-f4.tif
Fig. 4 (a) The packing arrangement of molecules viewed down the a-axis; (b) a plot of molecules of the title compound showing the formation of dimer by intermolecular C–H⋯O hydrogen bond.

3.3 Theoretical studies

3.3.1 Molecular geometry and QTAIM analysis. The optimized structure of the phenanthrenequinone 1 with labeled atoms is shown in Fig. 5. The corresponding value of optimized structural parameters (bond length, bond angle) calculated by B3LYP at 6-31+G(d,p) basis set are given in Table 4. The title compound contains methoxy groups and carbonyl functions attached with benzene nucleus. The calculated geometrical parameters show a good approximation and they are the bases for calculating other parameters, such as vibrational frequencies, thermodynamics and electronics properties. The benzene ring appears to be a little distorted and angles are slightly out of perfect hexagonal structure due to the substitution of the methoxy and carbonyl groups in the place of H-atom.
image file: c5ra21490d-f5.tif
Fig. 5 Optimized geometry of phenanthrenequinone 1 calculated at B3LYP/6-31+G(d,p) level.
Table 4 Selected bond lengths (angstroms) and bond angles (degrees) for phenanthrenequinone 1 calculated at the B3LYP/6-311++G(d,p) level
Parameter Calculated Parameter Calculated Parameter Calculated
C1–C2 1.425 C1–C6 1.371 C1–H13 1.086
C2–C3 1.440 C2–C15 1.416 C3–C4 1.433
C3–C7 1.442 C4–C5 1.397 C4–C8 1.495
C5–C6 1.415 C5–C11 1.499 C6–H14 1.084
C7–C17 1.369 C7–O22 1.349 C8–C9 1.512
C8–O20 1.219 C9–C10 1.354 C9–O21 1.338
C10–C11 1.464 C10–H12 1.083 C11–O19 1.232
C15–C16 1.387 C15–H18 1.085 C16–C17 1.398
C16–O23 1.356 O21–C24 1.429 O22–C28 1.427
O23–C32 1.429 C24–H25 1.095 C24–H26 1.089
C24–H27 1.095 C28–H29 1.097 C28–H30 1.095
C28–H31 1.090 C32–H33 1.090 C32–H34 1.096
C32–H35 1.096 C2–C1–C6 121.5 C2–C1–H13 118.3
C6–C1–H13 120.2 C1–C2–C3 118.5 C1–C2–C15 119.8
C3–C2–C15 121.6 C2–C3–C4 118.8 C2–C3–C7 116.6
C4–C3–C7 124.4 C3–C4–C5 119.7 C3–C4–C8 122.8
C5–C4–C8 116.6 C4–C5–C6 120.6 C4–C5–C11 121.2
C6–C5–C11 118.2 C1–C6–C5 120.1 C1–C6–H14 121.7
C5–C6–H14 118.1 C3–C7–C17 117.3 C3–C7–O22 118.9
C17–C7–O22 123.6 C4–C8–C9 116.9 C4–C8–O20 123.4
C9–C8–O20 119.3 C8–C9–C10 120.8 C8–C9–O21 112.1
C10–C9–O21 126.9 C9–C10–C11 120.5 C9–C10–H12 123.5
C11–C10–H12 115.9 C5–C11–C10 118.5 C5–C11–O19 120.6
C10–C11–O19 120.8 C2–C15–C16 120.3 C2–C15–H18 120.4
C16–C15–H18 119.3 C15–C16–C17 116.5 C15–C16–O23 119.3
C17–C16–O23 124.1 C7–C17–C16 126.6 C9–O21–C24 118.1
C7–O22–C28 117.0 C16–O23–C32 116.9 O21–C24–H25 110.8
O21–C24–H26 105.6 O21–C24–H27 110.6 H25–C24–C26 109.9
H25–C24–H27 109.8 H26–C24–H27 110.0 O22–C28–H29 110.8
O22–C28–H30 110.5 O22–C28–H31 105.8 H29–C28–H30 109.7
H29–C28–H31 109.9 H30–C28–H31 109.9 O23–C32–H33 105.8
O23–C32–H34 110.9 O23–C32–H35 110.9 H33–C32–H34 109.8
H33–C32–H35 109.7 H34–C32–H35 109.6 C7–C35–H36 119.7


The bonding within the AIM formalism is the existence of a bond path between two atoms and a bond critical point (BCP) in the middle of the path and it has been successfully applied in many previous studies.15,24 QTAIM analysis is based on quantum theory of atoms in molecule.25 This theory suggests that the values of some topological parameters at bond critical point (BCP) decide the nature of chemical interaction. Our AIM calculations on the dione reveal an intra-molecular O19⋯O21 interaction as depicted in molecular graph shown in Fig. 6. The bond-distance O19⋯O21 calculated at B3LYP/6-31+G(d,p) level is 2.643 Å. The values of topological parameters at BCP of O19⋯O21 are follows: charge density (ρ) = 0.016446 a.u., Laplacian (∇2ρ) = 0.059157 a.u., potential energy density (V) = −0.013225 a.u., kinetic energy density (G) = 0.0140073 a.u. and total energy density H (=V + G) = 0.000782 a.u. According to Rozas et al., this interaction should be characterized as weak and mainly electrostatic in nature for ∇2ρ > 0 and H > 0.26 Espinosa et al. have proposed a direct relationship between the interaction energy (ΔE) and V as, ΔE = −1/2V,27 therefore ΔE of O19⋯O21 interaction is estimated to be 4.147 kcal mol−1.


image file: c5ra21490d-f6.tif
Fig. 6 Molecular graph of phenanthrenequinone 1 showing intramolecular interaction (dotted line). Green and red points represent BCPs and RCPs respectively.
3.3.2 Vibrational analysis. In order to obtain a more complete description of the molecular motion of the phenanthrenequinone 1, the vibrational assignments of the calculated frequency are aided by the animation option of Gauss-view program, which gives a visual representation of the vibrational modes.28 The potential energy distribution is calculated with the help of VEDA 4 software program.29 Calculated frequencies are uniformly scaled with the factor of 0.9648 as devised by Merrick et al.30 to compensate the errors due to the neglect of anharmonic terms by the present theoretical model. The FTIR spectrum of these molecules are recorded using Shimadzu spectrometer in the region 4000–400 cm−1 using samples in KBr disc. Fig. 7 plots the simulated spectra along with FTIR spectra of molecules for a visual comparison.
image file: c5ra21490d-f7.tif
Fig. 7 Simulated and experimental FTIR spectra of phenanthrenequinone 1.

The experimental and calculated frequencies and the detailed description of each normal mode of vibration carried out in terms of their contribution to the potential energy are given along with corresponding FTIR values in Table 5.

Table 5 Vibrational analysis of prominent modes of for phenanthrenequinone 1 at B3LYP/6-311++G(d,p) levela
Calculated freq. (cm−1) Scaled freq. (cm−1) Intensity I.R FTIR freq. Assignment
a Abbreviations used. νas: asymmetric stretching; νs: symmetric stretching; σ: scissoring; ρ: rocking, τo: out of plane torrision; τi: in the plane torrision; R: carbon ring. Note: the PED distribution less than 15% are neglected for the sake of calculation.
3228 3114 4.2   R[νas(C10–H12)(99)]
3224 3110 5.2   R[νas(C6–H14)(100)]
3200 3087 5.0   R[νas(C15–H18)(100)]
3188 3076 7.3 3076 R[νas(C1–H13)(91)]
3169 3057 13.9   νas(C24–H25)(99)
3162 3051 12.7   νas(C28–H30)(92)
3159 3048 25.2   νas(C32–H34)(100)
3105 2996 24.5   νs(C24–H25)(99)
3094 2985 43.3   νs(C32–H34)(91)
3092 2983 12.7   νs(C28–H30)(89)
3034 2927 45.3 2945 νas(C24–H27)(100)
3029 2922 80.3   νs(C32–H35)(85)
3024 2917 32.2   νs(C28–H29)(97)
1752 1690 167.4 1670 [νas(C8–O20)(89)]
1708 1648 307.8 1648 [νas(C11–O19)(84)]
1665 1606 256.3 1620 R[νas(C9–C10)(62)]
1637 1579 231.5   R[νas(C1–C6)(42)]
1615 1558 114.8 1552 R[νas(C7–C17)(18) + νas(C1–C6)(15)]
1588 1532 8.8   R[νas(C2–C1)(22) + ρ(C2–C1–C6)(14)]
1516 1463 133.6 1475 R[νs(C2–C1)(19)]
1504 1451 26.9 1456 σ(H25–C24–H27)(58)
1503 1450 72.7   σ(H34–C32–H33)(30)
1498 1445 13.2   σ(H30–C28–H29)(34)
1497 1444 10.5   τ(H25–C24–H27)(77) + R[τo(H32–C31–C22–C14)(21)]
1491 1438 16.9   σ(H34–C32–H33)(30) + σ(H29–C28–H31)(19)
1490 1438 3.1   σ(H34–C32–H33)(23) + σ(H29–C28–H31)(24)
1488 1436 5.4   σ(H29–C28–H31)(71) + R[τi(H29–C28–O22–C7)(18)]
1485 1433 42.6   σ(H29–C28–H31)(33)
1478 1426 3.9   ω(H25–C24–H27)(79)
1452 1401 141.6   σ(H14–C6–C1)(11)
1420 1370 142.1 1363 R[νs(C7–C17)(25)] + σ(H29–C28–H31)(17)
1408 1358 25.2   R[νs(C1–C6)(26)]
1387 1338 25.1   R[νas(C2–C1)(15)]
1372 1324 89.7   R[σ(H12–C10–C9)(12)]
1344 1297 125.6   R[νas(C9–O21)(12)]
1330 1283 126.4   R[νs(C16–O23)(17) + νs(C3–C4)(15)]
1317 1271 199.3   R[νs(C9–O21)(15)]
1270 1225 229.5 1220 R[νas(C9–O21)(21) + σ(H12–C10–C11)(19)]
1231 1188 180.7   R[τi(H29–C28–O22–C7)(30) + τi(H34–C32–O23–C16)(22)]
1211 1168 142.0 1178 R[σ(H12–C10–C11)(27)] + τi(H25–C24–O21–C9)(16)
1206 1163 11.2   R[σ(H12–C10–C11)(14) + τo(H34–C32–O23–C16)(23)]
1198 1156 122.7   R[τi(H34–C32–O23–C16)(16)]
1187 1145 26.0   σ(H14–C6–C5)(44)
1172 1131 12.5   σ(H12–C10–C11)(20)
1171 1130 5.2   τ(H25–C24–H27)(23) + τi(H25–C24–O21–C9)(63)
1169 1128 1.5   R[τi(H29–C28–O22–C7)(34) + τi(H34–C32–O23–C16)(40)]
1167 1126 1.3   τ(H29–C28–H31)(17) + R[τi(H29–C28–O22–C7)(30)]
1146 1106 103.8 1105 R[νas(C3–C4)(15)]
1124 1084 198.4 1079 R[νas(C24–O21)(20)]
1081 1043 65.3   R[νs(C32–O23)(41)]
1037 1000 34.6   R[νs(C24–O21)(45)]
1003 970 4.6   R[τo(H12–C10–C11–C5)(63)]
999 964 25.9 954 R[νs(C24–O21)(39)]
978 944 26.0   R[νas(C16–O23)(16) + σ(C1–C6–C5)(25)]
928 895 11.5   R[σ(C6–C5–C4)(13)]
872 841 6.8 840 R[τo(H12–C10–C11–C5)(71)]
857 827 76.9   R[τi(H12–C10–C11–C5)(75)]
842 812 17.7   R[σ(C7–C17–C16)(12)]
816 787 2.6   R[τo(H12–C10–C11–C5)(20) + τo(C19–C5–C10–C11)(18)]
805 777 5.8   R[τi(H12–C10–C11–C5)(33)]
794 766 1.3   R[σ(C2–C1–C6)(22)]
757 730 2.8   R[τo(O21–C8–C10–C9)(20) + τo(C19–C5–C10–C11)(22)]
734 708 2.9 703 R[τo(C19–C5–C10–C11)(36)]
705 680 2.5   R[νs(C8–C4)(13)]
698 673 2.5   R[σ(O19–C11–C10)(30)]
653 630 12.2 640 R[σ(O23–C16–C17)(14)]
637 615 9.0   R[τo(O21–C8–C10–C9)(39)]
600 579 3.8   R[νas(C8–C4)(14)]
591 570 12.4 568 R[τo(C7–C17–C16–C15)(15) + τo(O23–C15–C17–C16)(23)]
570 550 3.7   R[τo(O23–C15–C17–C16)(23)]
552 532 1.4   R[νs(C3–C4)(10)]
519 501 7.0   R[σ(C7–C17–C16)(26)]
512 494 5.9 491 R[τo(O21–C8–C10–C9)(19)]
496 479 0.8   R[σ(C7–C17–C16)(17)]
486 467 10.7   R[ρ(C8–C4–C3)(12)]
434 419 1.0   R[σ(C32–O23–C16)(19)]



3.3.2.1 C–H vibration. The aromatic structure of the compound show the presence of C–H stretching vibration in the region 3141–2917 cm−1 which is the characteristic region (3100–3000 cm−1) for the identification of C–H stretching vibration.31 The C–H in plane bending frequencies appeared in the range 1300–1000 cm−1 are very useful for characterization purpose32 for title compound. The C–H out of plane bending vibrations are strongly coupled with vibrations occurring in the region 1000–750 cm−1.33 In FTIR the modes associated to C–H stretching are observed at 3076 cm−1 and 2945 cm−1. Other modes of vibration are observed in FTIR frequency at 1456 cm−1 and 1178 cm−1. The theoretically computed values for C–H vibrational modes by B3LYP/6-31+G(d,p) method are in excellent agreement with the experimental data.
3.3.2.2 C–O and C[double bond, length as m-dash]O vibrations. The carbonyl group vibrational frequencies are the significant characteristic bands in the vibrational spectra in ketones, and for this reason, such band have been the subject of extensive studies.34 The strong absorption band at 1648 and 1606 cm−1 in IR is ascribed to the stretching of C[double bond, length as m-dash]O group, lie in lower frequency region in the present case due to the delocalization of lone pair of electron. The observation are in good agreement with the value reported in the literature.35 The C[double bond, length as m-dash]O in-plane and out-of-plane bending mode is also assigned within the characteristic region and presented in Table 5. The bands observed at 1084, 1043, 1000, and 964 cm−1 are assigned to C–O stretching vibration and in FTIR at 1220 and 954 cm−1, which are further supported by their TED contribution. The scissoring and torsion modes of vibration are obtained at 673, 630 and below 600 cm−1 and assigned to the observed peaks at 640 and 491 cm−1 in the FTIR spectra.
3.3.2.3 C–C vibration. The C–C aromatic stretching vibrations give rise to characteristics bands in the spectral range from 1600–1400 cm−1. Therefore, the C–C stretching vibrations of phenanthrenequinone 1 are found in the region 1606–1463 cm−1 and corresponding FTIR spectra at 1620, 1552, 1475, 1363 and 1105 cm−1. Most of the ring vibrational modes are affected by the substitution in the aromatic ring. Other modes of vibrations such as scissoring, torsion are found at lower frequencies and their FTIR frequencies also listed in the Table 5.
3.3.3 NMR analysis. The combined use of NMR analysis and simulation methods provides a useful approach for the structural prediction of large biomolecules.36 The proton (1H) and carbon (13C) NMR chemical shift are calculated with GIAO (gauge including atomic orbital) method by applying B3LYP/6-31+G(d,p) method and compared with the experimental NMR spectra. Chemical shift of any proton ‘x’ is equal to the difference between isotropic magnetic shielding (IMS) of tetramethylsilane (TMS) and proton (x). It is defined by the equation: CSx = IMSTMS − IMSx. The experimental and calculated 1H and 13C values for the phenanthrenequinone 1 are presented in Table 6. The 1H and 13C chemical shifts are calculated in gas phase and CDCl3 solvent and are compared with experimental values, which are also measured in CDCl3 solvent.
Table 6 Selected NMR chemical shifts (δ, in ppm) of phenanthrenequinone 1 and their assignments
Atom δ Assignment Atom δ Assignment
Calc. Expt. Calc. Expt.
H12 5.84 6.00 [s, H(R1)] C2 138.97 138.95 [s, C(R2)]
H13 7.95 7.87 [s, H(R2)] C11 181.96 181.74 [s, C(R1)]
H14 8.20 8.04 [s, H(R2)] C15 102.21 101.92 [s, C(R3)]
H17 7.15 6.76 [s, H(R3)] C16 158.27 158.05 [s, C(R3)]
H33 3.92 3.94 [s, H(–OCH3)] C23 56.22 56.48 [s, C(–OCH3)]
H36 6.74 6.69 [d, H(R3)] C31 55.36 55.54 [s, C(–OCH3)]


The 1H chemical shift obtained and calculated for the hydrogen atoms of methoxy group(s) is relatively lower compared to the other hydrogen atoms within the molecule, and is quite expected as they are more shielded. Since the electron donating atom or group increases the shielding and moves the resonance towards to a lower frequency. The observed chemical shift values for the ring hydrogen are in good agreement with the experimental values and are presented in Table 6. The H12 atom chemical shift is slightly smaller than the H13, H14, H17 and H36. This is due to the electron donating oxygen atom that causes shielding of the aromatic proton. In 13C NMR spectrum, different signals are observed which are consistent with structure on the basis of molecular symmetry. Aromatic carbons show their chemical shift values between 100 and 150 ppm.37 From Table 6, it is clear that the calculated chemical shifts are in good agreement with the experimental chemical shifts. The atom C23 and C31 show lower values then the other carbon atoms presumably due to the shielding effect.

3.3.4 HOMO–LUMO and MESP surfaces analysis. The Highest Occupied Molecular Orbitals (HOMOs) and Lowest Unoccupied Molecular Orbitals (LUMOs) offer a reasonable qualitative prediction of the excitation properties and the ability of electron transportation.38 The HOMO primarily acts as an electron donor and LUMO acts as an electron acceptor and are the main orbitals which take part in chemical stability.39 The frontier molecular orbital gap has been used as a measure for the bioactivity, since the intra-molecular charge transfer determining the bioactivity depends on this energy gap. The HOMO and LUMO plots of the title compound are shown in Fig. 8. The energies of frontier molecular orbital of title compound are calculated by using B3LYP/6-31+G(d,p) method. The energy gap between Frontier molecular orbital of title compound is 2.869 eV. This value is smaller than that of a previously known natural phenanthrenequinone derivative, 5-hydroxy-3,6,7-trimethoxyphenanthrene-1,4-dione,16b (3.028 eV) calculated at the same level of theory. Thus, the smaller energy gap of phenanthrenequinone 1 suggests its enhanced chemical reactivity.
image file: c5ra21490d-f8.tif
Fig. 8 HOMO, LUMO and MESP surfaces of phenanthrenequinone 1 calculated at B3LYP/6-31+G(d,p) level.

The molecular electrostatic potential (MESP) have been used to predict the behavior and reactivity of the molecule. It is very useful in understanding the potential sites for electrophilic (negative region) and nucleophilic (positive region) reactions.40 MESP is also well suited for analyzing process based on the “recognition” of one molecule by another, as in drug receptor, and enzyme–substrate interactions, because it is through their potentials that the two species first “see” each other.41 To predict reactive sites for electrophilic and nucleophilic attack for the investigated molecule, MESP is calculated at the B3LYP/6-31+G(d,p) optimized geometries and shown in Fig. 8. The different values of the electrostatic potential at the surface are represented by different colors and potential increases in the order red < orange < yellow < green < blue. The color code of these maps is in the range between −0.04315 a.u. (deepest red) and 0.04315 a.u. (deepest blue) in the compound, where blue indicates the most electropositive i.e. electron poor region and red indicates the most electronegative region, i.e. electron rich region. From the MESP figure it is evident that the most electronegative region is located over oxygen atom attached by carbon atom which effectively acts as electron donor in molecule.

3.3.5 Electronic and thermodynamic parameters. Several electronic parameters viz. ionization potential (I), electron affinity (A), absolute electronegativity (χ) and chemical hardness (η) etc. are calculated at B3LYP/6-31+G(d,p) level. The electronic parameters of the molecule are calculated from the total energy and the Koopmans' theorem. The popular Koopmans' theorem describes ionization potentials (I) and electron affinity (A) as the negative of energy Eigen values of HOMO and LUMO respectively. Other parameters η and χ can be calculated by using finite-difference approximations42 as χ = 1/2(I + A) and η = 1/2(IA). Dipole moment (μ) of the molecule gives a signature about charge distribution and geometry of the molecule. These parameters are often used to describe chemical reactivity of molecule. The electronic parameters of title compound are listed in Table 7.
Table 7 Electronic and Thermodynamic parameters of phenanthrenequinone 1 calculated at B3LYP/6-31+G(d,p) level
Electronic parameter Dione Thermodynamic parameter Dione
I (eV) 5.782 ZPE (kcal mol−1) 173.44
A (eV) 2.913 E (kcal mol−1) 185.63
Eg (eV) 2.869 Cv (cal mol−1 K−1) 73.81
χ (eV) 4.348 S (cal mol−1 K−1) 142.85
η (eV) 2.869 H (kcal mol−1) 186.21
μ (a.u.) 4.957 G (kcal mol−1) 143.62


The thermodynamic parameters viz. zero point energy (ZPE), thermal energy at room temperature (E), heat capacity (Cv) and entropy (S) for title compound are also calculated and listed in Table 7. These parameters can be very useful in estimating reaction paths of molecules.

3.3.6 Prediction of cytotoxic property of phenanthrenequinone 1 from comparing structurally similar analogous compounds and molecular docking studies. Phenanthrenes are relatively uncommon class of secondary metabolites probably formed in plants by oxidative coupling of aromatic rings of stilbene precursors. They are mainly found in higher plants belonging to the Orchidaceae family although few of them have been reported to occur in the Hepaticae class and Dioscoreaceae, Combretaceae and Betulaceae families.17 Most natural 1,4-phenanthrenequinones occur in monomeric form and their structural diversity is mainly associated with the position of their oxygen functions (–OH and –OCH3) at C-2, C-3, C-5, C-6, C-7 and C-8 on 9,10-dehydro or dihydro skeletons and most of these plant derived 1,4-phenanthrenequinones are known to exhibit significant cytotoxic activity against different cancer cell lines.16b,17c,43a,b In prior investigations, several 1,4-phenanthrenequinones, viz., denbinodin[5-hydroxy-2,7-dimethoxy-1,4-phenanthrenequinone], sphenone A [3,6,7-trimethoxy-1,4-phenanthrenequinone], annoquinone A [3-methoxy-1,4-phenanthrenequinone], calanquinone B [6-hydroxy-3,5,7-trimethoxy-1,4-phenanthrenequinone] and 5-hydroxy-3,6,7-trimethoxy-1,4-phenanthrenequinone etc. are known to exhibit in vitro cytotoxic activity against various types of human cancer cell lines.17,43 The latter compound, 5-hydroxy-3,6,7-trimethoxy-1,4-phenanthrenequinone, showed significant cytotoxic activity against human lung, prostate, colon, breast and nasopharyngeal cancer cell lines with greatest activity against breast cancer MCF-7 cells (EC50 < 0.02 μg mL−1).16b The structural resemblance of 1 with the above compounds prompted us to undertake the molecular docking study of the title compound.

The molecular docking is a process by which the binding of a molecule (ligand) with a receptor is explored. The molecule binding to a receptor inhibits its function and thus acts effectively as a drug.44 As evidenced from the experimental observations that phenanthrenequinones exhibit significant potency in vitro against several cell lines, with the greatest activity against breast cancer MCF-7 cells.16b Hence, we have performed molecular docking studies of phenanthrenequinone 1 into epidermal growth factor receptor (EGFR) that is highly expressed in case of breast cancer (MCF-7 cells) in order to elicit whether it has analogous binding mode to EGFR inhibitor. We have carried out such studies using SWISSDOCK webserver.46 In this program docking runs are performed as blind by covering the entire protein and not defining any specific region of the protein as bonding pocket to avoid the sampling bias. Output clusters are obtained after each run and the result showed that cluster 0 is having the best full fitness (FF) score. A greater negative FF score indicates a more favorable binding mode with a better fit.

The crystal structure of EGFR along with its inhibitor (erlotinib) has been collected from protein data bank with PDB ID of 1M17.45,47 The docking picture obtained from UCSF chimera software of compound into EGFR has been shown in Fig. 9 along with its inhibitor. One can see that phenanthrenequinone 1 binds with EGFR in a fashion similar to erlotinib and shows the occurrence of one hydrogen bond with 2.200 Å, which is slightly larger than 2.059 Å. Thus, Fig. 9 demonstrates the binding model of phenanthrenequinone 1 in the LEU binding site and the results of this molecular docking can support the postulation that our active compound may act on the same enzyme target where EGFR inhibitor acts. The FF values obtained for the compound and inhibitor are −2175.73 kcal mol−1 and −2156.20 kcal mol−1, respectively and corresponding binding affinities are −7.07 kcal mol−1 and −6.97 kcal mol−1. The higher FF score and binding affinity for phenanthrenequinone 1 against EGFR shows its effective binding fit. Therefore, the compound reported here, can also be used as an inhibitor for epidermal growth factor receptor.


image file: c5ra21490d-f9.tif
Fig. 9 Docking of phenanthrenequinone 1 (a) and erlotinib inhibitor (b) into epidermal growth factor receptor. Hydrogen bonds are shown in the circle.

4. Conclusions

In conclusion, we have isolated a novel, naturally occurring trimethoxy phenanthrenequinone derivative from the yams of Dioscorea prazeri, an Indian medicinal plant, and its structure has been unequivocally established as 3,5,7-trimethoxyphenanthrene-1,4-dione (1) on the basis of its detailed spectral and single crystal X-ray analyses. The crystal structure was solved by direct methods using single-crystal X-ray diffraction data collected at room temperature and refined by full-matrix least-squares procedures to a final R-value of 0.0559 for 2832 observed reflections. The compound crystallizes in the triclinic space group P[1 with combining macron] with the following unit-cell parameters: a = 7.7045(3), b = 8.5088(4), c = 16.3254(7) Å, α = 98.908(2)°, β = 96.316(2)°, γ = 103.939(2)° and Z = 2. The crystal structure of the phenanthrenequinone 1 consists of three fused rings, and all the three chlorine atoms in the chloroform solvent molecule are found disordered over two set of sites with occupancy ratio 0.525[thin space (1/6-em)]:[thin space (1/6-em)]0.475. To our knowledge, the application of single crystal X-ray analysis for structure elucidation of a plant derived phenanthrenequinone derivative is reported for the first time in the present communication.

Exhaustive theoretical studies on the molecular structure, vibrational spectra, HOMO, LUMO, MESP surfaces, reactivity descriptors and molecular docking of this plant-derived new natural molecule have also been performed. The calculated vibrational spectral data matches well with the experimental observations on the phenanthrenequinone molecule. DFT calculations have been carried out using B3LYP/6-31+G(d,p) method. The QTAIM analysis reveals an inter-molecular interaction between oxygen atoms and characterizes it as a weak interaction. A FT-IR spectrum of phenanthrenequinone was recorded and studied experimentally and theoretically and has been found to be in good agreement. NMR analysis of the title compound is also in good agreement with the experimental data. In addition, the reactivity of the molecule was realized using various local as well as global molecular descriptors. Moreover, the molecular docking result suggests that the compound might exhibit inhibitory activity against epidermal growth factor receptor as the binding mode of the compound is similar to that of erlotinib, a marketed anticancer drug. Further study along this direction is underway in our laboratory.

Acknowledgements

This paper is dedicated to Professor David W. Allen (Sheffield Hallam University, UK) on the occasion of his 74th birthday. GB is thankful to the CSIR (New Delhi) and DST (New Delhi) for financial supports.

References

  1. R. N. Chopra, S. L. Nayar and I. C. Chopra, Glossary of Indian Medicinal Plants, CSIR, New Delhi, 1956, p. 98 Search PubMed.
  2. (a) L. P. K. Roy, U. K. Som, P. K. Ghosh, C. P. Dutta and M. Biswas, J. Indian Chem. Soc., 1989, 66, 289 CAS; (b) M. Santour and T. Mujamoto, Planta Med., 2004, 70, 90 CrossRef PubMed; (c) H. C. Chaturvedi, M. Jain and N. R. Kidwai, Indian J. Exp. Biol., 2007, 45, 937 CAS.
  3. K. Rajaraman and S. Rangaswami, Indian J. Chem., Sect. B: Org. Chem. Incl. Med. Chem., 1982, 21, 832 Search PubMed.
  4. R. N. Chakravari, D. Chakravari and M. N. Mitra, J. Indian Chem. Soc., 1961, 38, 635 Search PubMed.
  5. M. Wij, K. Rajaraman and S. Rangaswami, Indian J. Chem., Sect. B: Org. Chem. Incl. Med. Chem., 1977, 15, 451–454 Search PubMed.
  6. M. Biswas, U. K. Som, P. K. Ghosh, C. P. Dutta and A. Banerji, Tetrahedron, 1988, 44, 4871 CrossRef CAS.
  7. K. Rajaraman and S. Rangaswami, Indian J. Chem., 1975, 13, 1137 CAS.
  8. (a) G. Brahmachari, S. Sarkar, R. Ghosh, S. Barman, N. C. Mandal, S. K. Jash, B. Banerjee and R. Roy, Org. Med. Chem. Lett., 2014, 4, 65,  DOI:10.1186/s13588-014-0018-6; (b) G. Brahmachari, N. C. Mandal, R. Roy, R. Ghosh, S. Barman, S. Sarkar, S. K. Jash and S. Mondal, Fitoterapia, 2013, 90, 104 CrossRef CAS PubMed; (c) G. Brahmachari, N. C. Mandal, S. K. Jash, R. Roy, L. C. Mandal, A. Mukhopadhyay, B. Behera, S. Majhi, A. Mondal and A. Gangopadhyay, Chem. Biodiversity, 2011, 8, 1139 CrossRef CAS PubMed; (d) G. Brahmachari, R. Roy, L. C. Mandal, P. P. Ghosh and D. Gorai, J. Chem. Res., 2011, 35, 656 Search PubMed; (e) G. Brahmachari, L. C. Mandal, D. Gorai, A. Mondal, S. Sarkar and S. Majhi, J. Chem. Res., 2011, 35, 678 CrossRef CAS; (f) G. Brahmachari, S. K. Jash, A. Gangopadhyay, S. Sarkar, S. Laskar and D. Gorai, Indian J. Chem., Sect. B: Org. Chem. Incl. Med. Chem., 2008, 47, 1898 Search PubMed; (g) G. Brahmachari, D. Gorai, D. Chatterjee, S. Mondal and B. Mistri, Indian J. Chem., Sect. B: Org. Chem. Incl. Med. Chem., 2004, 43, 219 Search PubMed; (h) G. Brahmachari, D. Gorai, A. Gangopadhyay, S. Mondal and D. Chatterjee, J. Chem. Res. (S), 2003,(6), 362 CrossRef CAS; (i) K. S. Mukherjee, D. Gorai, S. M. A. Sohel, D. Chatterjee, B. Mistri, B. Mukherjee and G. Brahmachri, Fitoterapia, 2003, 74, 188 CrossRef CAS PubMed; (j) G. Brahmachari, D. Gorai, S. M. A. Sohel, S. Mondal and B. Mistri, J. Chin. Chem. Soc., 2003, 50, 325 CrossRef CAS; (k) G. Brahmachari and D. Chatterjee, Fitoterapia, 2002, 73, 363 CrossRef CAS PubMed; (l) K. S. Mukherjee, G. Brahmachari, T. K. Manna and P. K. Mukherjee, Phytochemistry, 1998, 49, 2533 CrossRef CAS.
  9. W. L. F. Armarego and C. L. L. Chai, Purification of Laboratory Chemicals, Butterworth-Heinemann, London, 5th edn, 2003 Search PubMed.
  10. G. M. Sheldrick, Acta Crystallogr., Sect. A: Found. Crystallogr., 2008, 64, 112 CrossRef CAS PubMed.
  11. D. Riedel, International Tables for X-ray Crystallography, ed. J. C. Wilson, Tables 4.2.6.8 and 6.1.1.4, Kluwer, Dordrecht, 1992, vol. C Search PubMed.
  12. ESI..
  13. M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuj, M. Caricato, X. Li, H. P. Hratchian, A. F. Iamaylov, J. Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J. A. Montgomery Jr, J. E. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Lyengar, J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador, J. J. Dannenberg, S. Dapprich, A. D. Daniels, O. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski and D. J. Fox, Gaussian 09, Revision A.1, Gaussian Inc., Wallingford CT, 2009 Search PubMed.
  14. (a) A. D. Becke, J. Chem. Phys., 1993, 98, 5648 CrossRef CAS; (b) C. Lee, W. Yang and R. G. Parr, Phys. Rev. B: Condens. Matter Mater. Phys., 1988, 1337, 785 CrossRef.
  15. (a) A. K. Srivastava and N. Misra, Can. J. Chem., 2014, 92, 234 CrossRef CAS; (b) A. K. Srivastava, B. Narayana, B. K. Sarojini and N. Misra, Indian J. Phys., 2014, 88, 547 CrossRef CAS; (c) A. K. Srivastava, V. Baboo, B. Narayana, B. K. Sarojini and N. Misra, Indian J. Pure Appl. Phys., 2014, 52, 507 Search PubMed; (d) A. K. Srivastava, A. K. Pandey, S. K. Gangwar and N. Misra, J. At. Mol. Sci., 2014, 5, 279 Search PubMed; (e) A. K. Srivastava, A. K. Pandey, S. Jain and N. Misra, Spectrochim. Acta, Part A, 2015, 136, 682 CrossRef CAS PubMed.
  16. (a) C. Honda and M. Yamaki, Phytochemistry, 2000, 53, 987 CrossRef CAS PubMed; (b) C. L. Lee, F. R. Chang, M. H. Yen, D. Yu, Y. N. Liu, K. F. Bastow, S. L. Morris-Natschke, Y. C. Wu and K. H. Lee, J. Nat. Prod., 2009, 72, 210 CrossRef CAS PubMed; (c) M. Udaya Bhaskar, L. Jagan Mohan Rao, N. S. Prakasa Rao and P. R. Mohana Rao, J. Nat. Prod., 1991, 54, 386 CrossRef; (d) Y. Masuda, R. Suzuki, H. Sakagami, N. Umemura and Y. Shirataki, Chem. Pharm. Bull., 2012, 60, 1216 CrossRef CAS PubMed.
  17. (a) B. Sritularak, M. Anuwat and K. Likhitwitayawuid, J. Asian Nat. Prod. Res., 2011, 13, 251 CrossRef CAS PubMed; (b) T.-H. Lin, S.-J. Chang, C.-C. Chen, J.-P. Wang and L.-T. Tsao, J. Nat. Prod., 2001, 64, 1084 CrossRef CAS PubMed; (c) Y. Tezuka, H. Hirano, T. Kikuchi and G.-J. Xu, Chem. Pharm. Bull., 1991, 39, 593 CrossRef CAS; (d) A. Kovács, A. Vasas and J. Hohmann, Phytochemistry, 2008, 69, 1084 CrossRef PubMed.
  18. L. J. Farrugia, J. Appl. Crystallogr., 1997, 30, 565 CrossRef CAS.
  19. L. J. Farrugia, J. Appl. Crystallogr., 1999, 32, 837 CrossRef CAS.
  20. A. L. Spek, Acta Crystallogr., Sect. D: Biol. Crystallogr., 2009, 65, 148 CrossRef CAS PubMed.
  21. M. Nardelli, J. Appl. Crystallogr., 1995, 28, 659 CrossRef CAS.
  22. F. H. Allen, O. Kennard, D. G. Watson, L. Brammer, A. G. Open and R. Taylor, J. Chem. Soc., Perkin Trans. 2, 1987, S1 RSC.
  23. (a) J. Bernstein, R. E. Davis, L. Shimoni and N.-L. Chang, Angew. Chem., Int. Ed., 1995, 34, 1555 CrossRef CAS; (b) M. C. Etter, J. C. MacDonald and J. Bernstein, Acta Crystallogr., Sect. B: Struct. Sci., 1990, 46, 256 CrossRef.
  24. A. Kumar, A. K. Srivastava, S. Gangwar, N. Misra, A. Mondal and G. Brahmachari, J. Mol. Struct., 2015, 1096, 94 CrossRef CAS.
  25. R. F. W. Bader, Atoms in Molecules. A Quantum Theory, Oxford University Press, New York, 2nd edn, 1990 Search PubMed.
  26. I. Rozas, I. Alkorta and J. Elguero, J. Am. Chem. Soc., 2000, 122, 11154 CrossRef CAS.
  27. E. Espinosa, E. Molins and C. Lecomte, Chem. Phys. Lett., 1998, 285, 170 CrossRef CAS.
  28. R. Dennington, T. Keith and J. Millam, Gauss View, Version 5, Semichem Inc., Shawnee Mission KS, 2009 Search PubMed.
  29. (a) M. H. Jamroz, Vibrational Energy Distribution Analysis, VEDA 4 Program, Warsaw, Poland, 2004 Search PubMed; (b) M. H. Jamroz, Spectrochim. Acta, Part A, 2013, 114, 220 CrossRef CAS PubMed.
  30. J. P. Merrick, D. Moran and L. Radon, J. Phys. Chem. A, 2007, 111, 11683 CrossRef CAS PubMed.
  31. (a) M. Silverstein, G. Glayton Basseler and C. Morrill, Spectrometric Identification of Organic Compounds, Wiley, New York, 1991 Search PubMed; (b) M. Arirazhagan and J. Senthil Kumar, Spectrochim. Acta, Part A, 2011, 82, 228 CrossRef PubMed.
  32. V. Arjunan, S. Thillai Govindaraya, P. Ravindran and S. Mohan, Spectrochim. Acta, Part A, 2014, 120, 473 CrossRef CAS PubMed.
  33. N. Sundaraganesan, S. Ilakiamani and B. D. Joshua, Spectrochim. Acta, Part A, 2007, 467, 287 CrossRef PubMed.
  34. B. Smith, Infrared Spectral Interpretation: A Systematic Approach, CRC Press, Washington DC, 1999 Search PubMed.
  35. (a) V. Arjunan, S. Subramanian and S. Mohan, Spectrochim. Acta, Part A, 2013, 105, 176 CrossRef PubMed; (b) M. Mukherjee and T. N. Misra, J. Raman Spectrosc., 1996, 27, 595 CrossRef.
  36. T. Schlick, Molecular Modelling and Simulation: An Interdisciplinary Guide, Springer, New York, second edn, 2010, vol. 21 Search PubMed.
  37. (a) H. O. Kalinowski, S. Berger and S. Braun, Carbon-13 NMR spectroscopy, John Wiley & sons, Chichester, 1998 Search PubMed; (b) Carbon-13 chemical shifts in structural and Sterochemical Analysis, ed. K. Pihlaja and E. Kleinpeter, VCH Publishers, Deerfield Beach, 1994 Search PubMed.
  38. (a) M. Belletete, J. F. Morin, M. Leclere and G. Durocher, J. Phys. Chem. A, 2005, 109, 6953 CrossRef CAS PubMed; (b) D. Zhenminga, S. Hepinga, L. Yufanga, L. Dianshenga and L. Bob, Spectrochim. Acta, Part A, 2011, 78, 1143 CrossRef PubMed.
  39. D. F. V. Lewis, C. Loannides and D. V. Parke, Xenobiotica, 1994, 24, 401 CrossRef CAS PubMed.
  40. (a) E. Scrocco and J. Tomasi, Adv. Quantum Chem., 1978, 103, 115 CrossRef; (b) F. J. Luqul, J. M. Lopez and M. Orozco, Theor. Chem. Acc., 2000, 103, 343 CrossRef.
  41. (a) E. Scrocco and J. Tomasi, Top. Curr. Chem., 1973, 7, 95 Search PubMed; (b) Y. Li, Y. Liu, H. Wang, X. Xiong, P. Wei and F. Li, Molecules, 2013, 18, 877 CrossRef CAS PubMed.
  42. R. G. Parr and W. Yang, Density Functional Theory of Atoms and Molecules, Oxford University Press, New York & Oxford, 1989 Search PubMed.
  43. (a) C.-L. Lee, K. Nakagawa-Goto, D. Yu, Y.-N. Liu, K. F. Bastow, S. L. Morris-Natschke, F.-R. Chang, Y.-C. Wu and K.-H. Lee, Bioorg. Med. Chem. Lett., 2008, 18, 4275 CrossRef CAS PubMed; (b) Y. H. Lee, J. D. Park, N. I. Baek, S. I. Kim and B. Z. Ahn, Planta Med., 1995, 61, 178 CrossRef CAS PubMed; (c) C. T. Kuo, B. C. Chen, C. C. Yu, C. M. Weng and M. J. Hsu, J. Biomed. Sci., 2009, 16, 43 CrossRef PubMed; (d) C. R. Yang, J. H. Guh, C. M. Teng, C. C. Chen and P. H. Chen, Br. J. Pharmacol., 2009, 157, 1175 CrossRef CAS PubMed.
  44. V. Srivastava, A. Kumar, B. N. Mishra and M. I. Siddiqui, Bioinformation, 2008, 4, 180 CrossRef.
  45. (a) J. Fricker, Lancet Oncol., 2006, 7, 621 CrossRef PubMed; (b) S. Madhusudan and T. S. Ganesan, Clin. Biochem., 2004, 37, 618 CrossRef CAS PubMed; (c) G. S. Cockerill and K. E. Lackey, Top. Med. Chem., 2002, 2, 1001 CrossRef CAS.
  46. A. Grosdidier, V. Zoete and O. Michielin, Nucleic Acids Res., 2011, 39, W270 CrossRef CAS PubMed.
  47. Available from http://www.rcsb.org/pdb/explore/explore.do?structureId=1M17.

Footnote

CCDC 1049913. For crystallographic data in CIF or other electronic format see DOI: 10.1039/c5ra21490d

This journal is © The Royal Society of Chemistry 2016
Click here to see how this site uses Cookies. View our privacy policy here.