Skip to content
BY-NC-ND 3.0 license Open Access Published by De Gruyter (O) February 7, 2017

Crystal structure of 5,7,4′-trihydroxy-3,8,3′-trymethoxyflavone, C18H16O8

  • Iván Brito EMAIL logo , Mario L. Simirgiotis , Leyla Pasten , Rubén Muñoz , Jorge Bórquez and Alejandro Cárdenas

Abstract

C18H16O8, monoclinic, P21/n, (no. 14), a = 7.9253(3) Å, b = 16.8928(6) Å, c = 11.8439(5) Å, β = 92.387(2)°, V = 1584.29(11) Å3, Z = 4, Rgt(F) = 0.043, wRref(F2) = 0.102, T = 295(2) K.

CCDC no.:: 1526063

The asymmetric unit of the title crystal structure is shown in the figure. Tables 1 and 2 contain details on crystal structure and measurement conditions and a list of the atoms including atomic coordinates and displacement parameters.

Table 1

Data collection and handling.

Crystal:Yellow block
Size:0.30 × 0.22 × 0.14 mm
Wavelength:Mo Kα radiation (1.54178 Å)
μ:10.2 cm−1
Diffractometer, scan mode:Bruker AXS, φ and ω
2θmax, completeness:118.2°, >99%
N(hkl)measured, N(hkl)unique, Rint:32631, 2274, 0.089
Criterion for Iobs, N(hkl)gt:Iobs > 2 σ(Iobs), 1510
N(param)refined:242
Programs:SHELX [10], Bruker programs [11], OLEX2 [12] PLATON [13]
Table 2

Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2).

AtomxyzUiso*/Ueq
O10.5544(2)0.66230(9)0.31502(14)0.0388(5)
O20.4170(2)0.52051(9)0.25114(15)0.0443(5)
O30.0958(2)0.52054(10)0.14519(18)0.0597(6)
H30.14710.48310.17440.090*
O40.0675(2)0.79701(10)0.15748(18)0.0539(6)
H40.12090.83530.18200.081*
O50.3169(2)0.87250(10)0.26028(16)0.0478(5)
O60.6286(2)0.87039(9)0.37013(15)0.0422(5)
O71.1846(2)0.80685(10)0.52927(17)0.0554(6)
O81.2827(2)0.65493(10)0.52644(17)0.0531(6)
H81.33110.69410.55290.080*
C1’0.8013(3)0.71289(14)0.4003(2)0.0342(6)
C2’0.9090(3)0.77182(15)0.4446(2)0.0393(7)
H2’0.87210.82410.44570.047*
C3’1.0688(3)0.75358(15)0.4867(2)0.0385(7)
C4’1.1240(3)0.67539(15)0.4881(2)0.0378(7)
C5’1.0176(3)0.61722(15)0.4478(2)0.0476(8)
H5’1.05300.56470.45040.057*
C6’0.8589(3)0.63525(15)0.4034(2)0.0438(7)
H6’0.78930.59490.37520.053*
C20.6344(3)0.73097(14)0.3490(2)0.0343(6)
C30.5555(3)0.80169(14)0.3315(2)0.0343(6)
C40.3905(3)0.80718(14)0.2762(2)0.0367(7)
C50.1549(3)0.72954(15)0.1825(2)0.0376(7)
C60.0870(3)0.65860(15)0.1502(2)0.0437(7)
H6−0.01690.65690.11080.052*
C70.1727(3)0.58862(15)0.1761(2)0.0400(7)
C80.3295(3)0.59027(13)0.2321(2)0.0350(6)
C90.3981(3)0.66317(14)0.2619(2)0.0325(6)
C100.3141(3)0.73404(13)0.2397(2)0.0332(6)
C160.4285(4)0.49738(16)0.3677(2)0.0574(9)
H16B0.51300.45700.37820.086*
H16A0.32130.47720.38950.086*
H16C0.45880.54240.41360.086*
C170.7073(4)0.91556(16)0.2842(3)0.0623(9)
H17C0.62660.92590.22330.093*
H17A0.80060.88620.25630.093*
H17B0.74740.96480.31560.093*
C181.1472(4)0.88914(15)0.5160(3)0.0551(8)
H18C1.11560.89980.43840.083*
H18A1.24520.91980.53790.083*
H18B1.05580.90310.56280.083*

Source of material

The title compound C18H16O8 was isolated using high speed countercurrent chromatography (HSCCC). Following our program to isolate components from the Atacama Desert Flora, Northern Chile [1], [2], [3], [4], dried aerial parts of Parastrephia quadrangularis (1622 g) collected in april 2015 in El Tatio, were defatted with hexane (3 liters, 3 times in the dark, 24 h each time) and 54.82 g were obtained after evaporation of the solvent. Then the plant material was extracted with ethyl acetate (3 liters, 3 times in the dark, 24 Hs each time. After evaporation of the solvent under vacuo at 35 oC, 485 g of a dark gummy extract was obtained. A portion of the extract (0.5 g) was filtered and submitted to a HSCCC centrifuge (Quattro MK-7, Bridgend, UK) for the separation of its components. The solvent system selected was the isocratic two-phase non aqueous solvent system: n-hexane: ethyl acetate: methanol: water 3:7:5:5 v/v/v provided the better K values for all major compounds (0.5 < K < 1.3). This system was previously used for the separation of terpenes [4]. After equilibration of the solvent system in a separatory funnel, the upper and lower working phases were separated and degassed in an ultrasonic bath for 15 min before use. The sample was prepared by dissolving 500 mg of exudate from P. quadrangularis in 5.0 mL of each phase of the solvent system, filtered and loaded into an injection valve (Rheodyne model 5010A) equipped with a 8 mL loop. The preparative coil (116 mL) was filled with the upper stationary phase and the apparatus was rotated at 850 rpm. The mobile lower phase was then pumped in a head-to-tail direction (H-T) at a flow rate of 5 mL/minute. After the mobile phase front emerged and the hydrodynamic equilibrium was established in the column, the percentage of the retention of the stationary phase (60%) was recorded. Then the sample was injected thorough the injection valve at a flow rate of 5 mL-minute. The fractions eluted were collected with the fraction collector (5 mL each) and analyzed by TLC (F254 Silica gel plates, developed with hexane : EtOAc, 1 : 1 v/v, and spots visualized by spraying with vanillin : sulfuric acid 2% in ethanol and heating. CCC rotation was interrupted in tube 70 and the coil content was collected (“wash-off”), originating 95 fractions of 6 mL each. After re-purification by sephadex LH 20 (solvent methanol), HSCCC fractions 12–18 afforded 11-p-coumaroyl-tremetone [5], (55 mg) fractions 23–29 umbelliferone [6] (12 mg) and fractions 35–38 the title compound (5, 7, 4′ trihydroxy-3,8,3′ trymethoxyflavone, 12 mg) for which NMR data are consistent with literature [7], [8], [9]. Recrystallization from ethyl acetate at −20° C yielded yellow crystals (5 mg), m.p. 265–267 °C. The molecular weight was determined by orbitrap ESI-MS/MS with a mass spectrometer (Q-exactive Focus, Bremen, Germany) [M-H]: 359.0771, calcd. for C18H17O8: 359.0772. 1H-NMR (Bruker Avance 300 MHz, DMSO-d6) δ ppm: 7.65(1H, d, J = 1.4 Hz, H-2′), 7.55(1H, dd, J = 1.4, 8.4 Hz, H-6′), 6.97(1H, d, J = 8.4 Hz, H-5′), 6.43(1H, s, H-6), 3.90(3H, s, OCH3), 3.93(3H, s, OCH3), 3.86(3H, s, OCH3). 13C-NMR (Bruker Avance 300 MHz, DMSO-d6) δ ppm: 147.92 (C-2), 137.23 (C-3), 177.35 (C-4), 162.54 (C-5), 99.35 (C-6), 165.74 (C-7), 163.2 (C-8), 158.2 (C-9), 104.4 (C-10), 124.13 (C-1), 116.06 (C-2′), 146.23 (C-3′), 148.75 (C-4′), 116.23 (C-5′), 121.66 (C-6′), 56.92 (O-CH3), 56.90 (O-CH3), 56.95 (O-CH3).

Experimental details

H atoms were located in the difference Fourier map, but refined with fixed individual displacement parameters, using a riding mode with C—H distances of 0.93 Å (for aromatic rings), 0.96 Å (for CH3), with Uiso(H) values of 1.2Ueq(C) (for CH in aromatic), and 1.5Ueq(C) (for methyl group), O—H distances are 0.82 Å with Uiso(H) values of 1.5Ueq(O).

Discussion

The core system of the title compound is planar (r.m.s. deviation 0.072 Å), excluding the methyl groups C16 and C17. The molecular structure is stabilized by an intramolecular hydrogen bond, O4—H4⋯O5. In the crystal packing all hydroxyl groups participate as donor and acceptor of intermolecular hydrogen bonds interactions. In total different set-graphs motifs such as chains and rings are present [14]. The packing also features π–π stacking interactions between benzene rings [centroid–centroid distance = 3.701(2) Å]. All distances and angles are in the normal range.

Acknowledgements

The authors thank FONDECYT (Chile) (Grant 1140178) for financial support. IB Thanks to Fondequip (EQM13–0021).

References

1. Brito, I.; Simirgiotis, M. J.; Brito, A.; Werner, M. R.; Bórquez, J.; Winterhalter, P.; Cárdenas, A.: A non-centrosymmetric polymorph of 5-hydroxy-7-methoxy-2-phenylchroman-4-one. J. Chil. Chem. Soc. 60 (2015) 2864–2866.10.4067/S0717-97072015000100020Search in Google Scholar

2. Brito, I.; Bórquez, J.; Simirgiotis, M.; Neves-Vieira, M.; Jerz, G.; Winterhalter, P.; Bolte, M.; Cárdenas, A.: Crystal structure of 2-nor-1,2-secolycoserone, C24H32O 4. Z. Kristallogr. − NCS. 229 (2014) 399–400.10.1515/ncrs-2014-0212Search in Google Scholar

3. Bórquez, J.; Ardiles, A.; Loyola, L. A.; Peña-Rodriguez, L. M.; Molina-Salinas, G. M.; Vallejos, J.; Collado, I. G.; Simirgiotis, M. J.: Further mulinane and azorellane diterpenoids isolated from mulinum crassifolium and azorella compacta. Molecules 19 (2014) 3898–3908.10.3390/molecules19043898Search in Google Scholar

4. Bórquez, J.; Bartolucci, N. L.; Echiburú-Chau, C.; Winterhalter, P.; Vallejos, J.; Jerz, G.; Simirgiotis, M. J.: Isolation of cytotoxic diterpenoids from the Chilean medicinal plant Azorella compacta Phil from the Atacama Desert by high–speed counter–current chromatography. J. Sci. Food Agric. 96 (2016) 2832–2838.10.1002/jsfa.7451Search in Google Scholar

5. Bohlmann, F.; Fritz, U.; King, R. M.: Neue tremeton-derivate aus Parastrephia lepidophylla. Phytochemistry. 18 (1979) 1403–1405.10.1016/0031-9422(79)83037-XSearch in Google Scholar

6. Pan, L.; Li, X.-Z.; Yan, Z.-Q.; Guo, H.-R.; Qin, B.: Phytotoxicity of umbelliferone and its analogs: Structure–activity relationships and action mechanisms. Plant Physiol. Biochem. 97 (2015) 272–277.10.1016/j.plaphy.2015.10.020Search in Google Scholar PubMed

7. Yoo, H.; Kim, S. H.; Lee, J.; Kim, H. J.; Seo, S. H.; Chung, B. Y.; Jin, C.; Lee, Y. S.: Synthesis and antioxidant activity of 3-methoxyflavones. Bull. Korean Chem. Soc. 26 (2005) 2057–2060.10.1002/chin.200616142Search in Google Scholar

8. Agrawal, P. K.: Carbon-13 NMR of flavonoids. Elsevier: Michigan, 1989; p 564.10.1016/B978-0-444-87449-8.50011-0Search in Google Scholar

9. Modak, B.; Rojas, M.; Torres, R.: Chemical Analysis of the Resinous Exudate Isolated fromHeliotropium taltalenseand Evaluation of the Antioxidant Activity of the Phenolics Components and the Resin in Homogeneous and Heterogeneous Systems. Molecules 14 (2009) 1980–1989.10.3390/molecules14061980Search in Google Scholar PubMed PubMed Central

10. Sheldrick, G. M.: Crystal structure refinement with SHELXL. Acta Crystallogr. C71 (2015) 3–8.10.1107/S2053229614024218Search in Google Scholar PubMed PubMed Central

11. Bruker. APEX2, SAINT and SADABS. Bruker AXS Inc., Madison, Wisconsin, USA, 2007.Search in Google Scholar

12. Dolomanov, O. V.; Bourhis, L. J.; Gildea, R. J.; Howard, J. A. K.; Puschmann, H.: OLEX2: a complete structure solution, refinement and analysis program. J. Appl. Cryst. 42 (2009) 339–341.10.1107/S0021889808042726Search in Google Scholar

13. Spek, A. L.: Structure validation in chemical crystallography. Acta Crystallogr. D65 (2009) 148–155.10.1107/S090744490804362XSearch in Google Scholar PubMed PubMed Central

14. 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. 34 (1995) 1555–1573.10.1002/anie.199515551Search in Google Scholar

Received: 2016-7-22
Accepted: 2017-1-7
Published Online: 2017-2-7
Published in Print: 2017-3-1

©2017 Iván Brito et al., published by De Gruyter.

This work is licensed under the Creative Commons Attribution-NonCommercial-NoDerivatives 3.0 License.

Downloaded on 25.5.2024 from https://www.degruyter.com/document/doi/10.1515/ncrs-2016-0224/html
Scroll to top button