Acessibilidade / Reportar erro

Friedelane Triterpenes with Cytotoxic Activity from the Leaves of Maytenus quadrangulata (Celastraceae)

Abstract

Three new triterpenes, 3,4-seco-3,11β-epoxyfriedel-4(23)-en-3β-ol (1), friedelan-3α,11β diol (2), 7β,26-epoxyfriedelan-3a,7a-diol (3), a mixture of two new triterpenes 3α-hydroxyfriedelan-29-yl palmitate (4) and 3α-hydroxyfriedelan-29-yl stearate (5) and eleven known compounds were obtained from the hexane extract of Maytenus quadrangulata leaves. The structures and the relative stereochemistry of the new triterpenes were established through 1D/2D nuclear magnetic resonance (NMR), high resolution mass spectrometry (HRMS), and Fourier transform infrared (FTIR) spectral data. The hexane extract and isolated compounds were submitted to the cytotoxicity assays against leukemia (THP-1 and K562), ovarian (TOV-21G) and breast cancer (MDA-MB-231) cell lines. Compounds 1, 2 and 11β-hydroxyfriedelan-3-one (15) displayed high cytotoxicity and selectivity against leukemic cells when compared to positive control cytarabine (for THP-1) and imatinib (for K562). Furthermore, compound 2 showed similar cytotoxicity and an enhanced selectivity towards ovarian and breast cells in comparison to positive control etoposide.

Keywords:
Celastraceae; Maytenus; pentacyclic triterpenes; cytotoxicity


Introduction

The Celastraceae family is composed of 1210 species distributed in 96 genera, which are found mainly in tropical and subtropical regions of South America, Asia and North Africa.11 Simmons, M. P.; Cappa, J. J.; Archer, R. H.; Ford, A. J.; Eichstedt, D.; Clevinger, C. C.; Mol. Phylogenet. Evol. 2008, 48, 745.,22 Christenhusz, M. J. M.; Byng, J. W.; Phytotaxa 2016, 261, 201. The genus Maytenus belongs to the Celastraceae family and displays different species commonly used in folk medicine to treat stomach problems and due to its anticancer, analgesic, anti-ulcerogenic, antiasthmatic, anti-inflammatory, anti-human immunodeficiency virus (HIV) and antimicrobial properties.33 Mokoka, T. A.; McGaw, L. J.; Mdee, L. K.; Bagla, V. P.; Iwalewa, E. O.; Eloff, J. N.; BMC Complementary Altern. Med. 2013, 13, 111.,44 Zhang, L.; Ji, M.-Y.; Bin, Q.; Li, Q.-Y.; Zhang, K.-Y.; Liu, J.-C.; Dang, L.-S.; Li, M.-H.; Med. Chem. Res. 2020, 29, 575. Maytenus quadrangulata (Schrad.) Loes, popularly known as “espinho-de-deus”, is naturally found in the Atlantic Forest and Caatinga (savanna region) biome of Brazil in the states of Bahia, Espírito Santo and Minas Gerais.55 Cardoso, D. B. O. S.; de Queiroz, L. P.; J. Bot. Res. Inst. Texas 2008, 2, 551.

6 Menino, G. C. O.; dos Santos, R. M.; Apgaua, D. M. G.; Pires, G. G.; Pereira, D. G. S.; Fontes, M. A. L.; Almeida, H. S.; CERNE 2015, 21, 277.
-77 Celastraceae in Flora do Brasil 2020, Jardim Botânico do Rio de Janeiro, http://floradobrasil.jbrj.gov.br/reflora/floradobrasil/FB16751, accessed in March 2022.
http://floradobrasil.jbrj.gov.br/reflora...
Even though its great pharmacological potential, no phytochemical and biological evaluation involving M. quadrangulata was found in the literature.

Pentacyclic triterpenes are secondary metabolites commonly found in the genus Maytenus.44 Zhang, L.; Ji, M.-Y.; Bin, Q.; Li, Q.-Y.; Zhang, K.-Y.; Liu, J.-C.; Dang, L.-S.; Li, M.-H.; Med. Chem. Res. 2020, 29, 575.,88 Sousa, G.; Duarte, L.; Alcântara, A.; Silva, G.; Vieira-Filho, S.; Silva, R.; Oliveira, D.; Takahashi, J.; Sousa, G. F.; Duarte, L. P.; Alcântara, A. F. C.; Silva, G. D. F.; Vieira-Filho, S. A.; Silva, R. R.; Oliveira, D. M.; Takahashi, J. A.; Molecules 2012, 17, 13439. In recent years, such compounds have been increasingly studied due to their wide range of pharmacological activities, including antitumor, hepatoprotective, anti-inflammatory, hypoglycemic, antibacterial, antioxidant, anti-HIV, anti-trypanosome, leishmanicidal, antiarthritis and immunological adjuvant activities.99 Xu, C.; Wang, B.; Pu, Y.; Tao, J.; Zhang, T.; J. Sep. Sci. 2017, 41, 6. For example, friedelan-3-one exhibited gastric anti-ulcer, anti-inflammatory (in vivo), analgesic, antipyretic, and antioxidant activities.1010 Antonisamy, P.; Duraipandiyan, V.; Aravinthan, A.; Al-Dhabi, N. A.; Ignacimuthu, S.; Choi, K. C.; Kim, J.-H.; Eur. J. Pharmacol. 2015, 750, 167.

11 Antonisamy, P.; Duraipandiyan, V.; Ignacimuthu, S.; J. Pharm. Pharmacol. 2011, 63, 1070.
-1212 Sunil, C.; Duraipandiyan, V.; Ignacimuthu, S.; Al-Dhabi, N. A.; Food Chem. 2013, 139, 860. Furthermore, friedelan-3-one and some of its derivatives, obtained from reactions involving the ketone group at C-3, also demonstrated in vitro cytotoxic activity against THP-1 and K562 leukemia cell lines.1313 Aguilar, M. G.; Sousa, G. F.; Evangelista, F. C. G.; Sabino, A. P.; Vieira Filho, S. A.; Duarte, L. P.; Nat. Prod. Res. 2020, 34, 810.

In the present work, the phytochemical study of the M. quadrangulata hexane extract of the leaves led to the isolation of the three new triterpenes 3,4-seco-3,11β epoxyfriedel-4(23)-en-3β-ol (1), friedelan-3α,11β diol (2), 7β,26-epoxyfriedelan-3,7a-diol (3) and a mixture of the two new compounds 3α-hydroxyfriedelan-29-yl palmitate (4) and 3α-hydroxyfriedelan-29-yl stearate (5). Moreover, eleven known compounds were also obtained: friedelan-3-one (6), friedelan-3β-ol (7), friedelan-3α-ol (8), friedelan-3,7-dione (9), 3β-hydroxyfriedelan-7-one (10), 3α-hydroxyfriedelan-7-one (11), gutta-percha polymer (12), 3,4-seco-friedelan-3,11-olide (13), β-sitosterol (14), 11β-hydroxyfriedelan-3-one (15) and mixture of friedelan-3α,11β-diol (2) and friedelan-3β,11β diol (16) (Figure 1). The structural elucidation of the new triterpenes 1, 2, 3, 4 and 5 was accomplished by infrared (IR), 1H and 13C nuclear magnetic resonance (NMR), including two-dimensional experiments (heteronuclear single quantum coherence (HSQC), heteronuclear multiple bond correlation (HMBC), correlated spectroscopy (COSY) and nuclear overhauser effect spectroscopy (NOESY)), high-resolution atmospheric pressure chemical ionization mass spectrometry (HR-APCI-MS) and high-resolution electrospray ionization mass spectrometry (HR-ESI-MS) analyses. The cytotoxicity of the compounds 1, 2, 6-11, 13, 15 and the hexane extract was evaluated against leukemia (THP-1 and K562), ovarian (TOV-21G) and breast cancer (MDA-MB-231) cell lines.

Figure 1
Compounds 1-16 from Maytenus quadrangulata leaves.

Experimental

Column chromatography (CC) was carried out with silica gel 60 (230-400 mesh or 70-230 mesh, Merck, Darmstadt, Germany). Medium pressure liquid chromatography (MPLC) was performed on a Biotage Isolera Spektra One system (Biotage, Uppsala, Sweden), using a Biotage Snap cartridge of 50 g (Biotage, Uppsala, Sweden). The mobile phases were hexane, chloroform (CHCl3), dichloromethane (CH2Cl2), ethyl acetate (AcOEt) and methanol, pure or in mixtures of increasing polarity. All solvents were Vetec (Duque de Caxias, Brazil) or Sigma-Aldrich (Saint Louis, USA), analytical grade. Pre-coated plates with silica gel 60 G (Vetec, Duque de Caxias, Brazil) were used in all thin layer chromatography (TLC) procedures. The melting temperatures of isolated compounds were determined on a Microchemical apparatus MQAPF-302 (Microchemical, Palhoça, Brazil), without correction for normal temperature and pressure conditions. The infrared spectra were obtained on a spectrometer Spectrum One PerkinElmer attenuated total reflection (ATR) (PerkinElmer, Massachusetts, USA) or Shimadzu IR-408 (Shimadzu, Singapore) (sample in KBr disk). 1D/2D NMR spectra were recorded on Bruker Avance DRX-400 and DRX-200 spectrometers (Bruker Avance, Billerica, USA), using CDCl3 or CDCl3 with some drops of pyridine (Py-d5), as solvent. The chemical shifts (δ) were recorded in ppm using tetramethylsilane (TMS) as internal standard and the coupling constants (J) are given in Hz. The mass spectra were acquired in a high-resolution mass spectrometer of the Thermo Scientific Q Exactive Orbitrap type (Thermo Fisher Scientific, Bremen, Germany), in positive or negative mode. The operating parameters of APCI ionization source were: spray voltage 5000 V, flow of 20 µL min-1, sheath gas 10, capillary temperature 320 °C, auxiliary gas temperature 37 °C, s-lens 50. The operating parameters of HR-ESI ionization source were: spray voltage 5200 V, flow of 20 µL min-1, sheath gas 12, capillary temperature 300 °C, auxiliary gas temperature 37 °C, s-lens 50. The compounds were detected and confirmed by the exact mass and isotopic parameters obtained in MS11 Simmons, M. P.; Cappa, J. J.; Archer, R. H.; Ford, A. J.; Eichstedt, D.; Clevinger, C. C.; Mol. Phylogenet. Evol. 2008, 48, 745. and MS22 Christenhusz, M. J. M.; Byng, J. W.; Phytotaxa 2016, 261, 201. spectra.

Plant material

Leaves of Maytenus quadrangulata were collected in the district of Brejo do Amparo, municipality of Januária (44º24’17.48’’ W and 15º25’10.85’’ S), Minas Gerais, Brazil, in February 2017. A voucher specimen (HMC, No. 405) was deposited in the Herbarium of the Departamento de Botânica, Instituto de Ciências Biológicas, Universidade Federal de Minas Gerais. The botanical material was registered at Conselho de Gestão do Patrimônio Genético (CGEN/SisGen), Brazil, under the number AAECF38.

Extraction and isolation

Leaves of M. quadrangulata were dried at room temperature and ground then, the powder (1743.0 g) was subjected to exhaustive extraction by maceration with hexane. A solid was formed during the partial removal of the hexane using a rotary evaporator (solid of the hexane extract (SHE), 465.0 mg, 1.3% of the hexane extract (HE)) and was filtered under reduced pressure. After filtration, the rest of the hexane has been completely removed to afford the hexane extract (HE: 36.3 g).

SHE (450.0 mg,) was subjected to CC using silica gel (50.8 g, 70-230 mesh) yielding 113 fractions, which were gathered in 8 groups (A1-A8), according to the TLC profiles. From these groups were obtained: friedelan-3-one (6) (A1 16.1 mg, hexane-CH2Cl2 7:3 v/v, 3.6% of SHE), mixture of compounds 6 and friedelan-3β-ol (7) (A2 82.2 mg, hexane-CH2Cl2 7:3 v/v, 18.3% of SHE), mixture of compounds 6, 7 and friedelan-3α-ol (8) (A3 30.9 mg, hexane-CH2Cl2 6:4 v/v, 6.9% of SHE), friedelan-3α-ol (8) (A4-87.4 mg, hexane-CH2Cl2 6:4 v/v, 19.4% of SHE), friedelan-3,7-dione (9) (A5-6.9 mg, hexane-CH2Cl2 1:1 v/v, 1.5% of SHE), mixture of 3α-hydroxyfriedelan-29 yl palmitate (4) and 3α-hydroxyfriedelan-29-yl stearate (5) (A6-5.0 mg, hexane-CH2Cl2 3:7 v/v, 1.1% of SHE) and 3β-hydroxyfriedelan-7-one (10) (A7-11.0 mg, CH2Cl2-AcOEt 1:1 v/v, 2.4% of SHE). The group A8 was chromatographed (silica gel 230-400 mesh) to give 3α-hydroxyfriedelan-7-one (11) (14.0 mg, CH2Cl2-AcOEt 95:05 v/v, 3.1% of SHE).

The hexane extract (HE, 35.3 g) was subjected to CC using silica gel (765.7 g, 70-230 mesh) obtaining 190 fractions, which were gathered in 16 groups (B1 B16). From the groups B5 and B6 were obtained the gutta-percha polymer (12) (218.3 mg, hexane-CH2Cl2 7:3 v/v, 0.6% of HE) and mixture of compounds 6 and 7 (286.8 mg, hexane-CH2Cl2 7:3 v/v, 0.8% of HE), respectively. The other groups were submitted to successive CC (230 400 mesh), providing the constituents: compound 7 (B7-90.7 mg, hexane-CH2Cl2 6:4 v/v, 0.3% of HE), 3,4-seco-friedelan-3,11-olide (13) (B7-444.3 mg, hexane-CH2Cl2 1:9 v/v and B8-658.5 mg, hexane-CH2Cl2 1:1 v/v, 3.1% of HE), compound 8 (B9 14.5 mg, hexane-CH2Cl2 1:1 v/v, 0.04% of HE), 3,4-seco-3,11β-epoxyfriedel-4(23)-en-3β-ol (1) (B9 39.6 mg, hexane-CH2Cl2 1:1 v/v, 0.1% of HE), β-sitosterol (14) (B10-158.5 mg, hexane-CH2Cl2 4:6 v/v, 0.4% of HE), mixture of compounds 4 and 5 (B11 37.2 mg, hexane-CH2Cl2 4:6 v/v, 0.1% of HE), compound 10 (B12-14.9 mg, hexane-CH2Cl2 4:6 v/v, 0.04% of HE), 11β-hydroxyfriedelan-3-one (15) (B13 23.3 mg, CH2Cl2 AcOEt 9:1 v/v and B14 57.2 mg, CH2Cl2 AcOEt 9:1 v/v, 0.2% of HE), the mixture of friedelan-3α,11β diol (2) and friedelan-3β,11β-diol (16) (B15-19.0 mg, CH2Cl2 AcOEt 8:2 v/v, 0.05% of HE), friedelan-3α,11β diol (2) (B15-198.7 mg, CH2Cl2-AcOEt 8:2 v/v, 0.6% of HE), and 7β,26-epoxyfriedelan-3a,7a-diol (3) (B16-29.7 mg, CH2Cl2-AcOEt 6:4 v/v, 0.08% of HE).

3,4-Seco-3,11β-epoxyfriedel-4(23)-en-3β-ol (1)

White solid; mp 121-123 °C; IR (ATR) ν / cm-1 3436, 3224, 3006, 2943, 2931, 1637, 1453, 1384, 1115, 1035, 1026, 994, 907; 1H and 13C NMR data see Table 1; HRMS (APCI) (positive-ion mode) m/z, calcd. for C30H51O2+ [M + H]+: 443.3884, found: 443.3886.

Table 1
1H (400 MHz) and 13C NMR (100 MHz) data of compounds 1, 2, 3, 4, 5 and 11 isolated from hexane extract of Maytenus quadrangulata leaves

Friedelan-3α,11β-diol (2)

White solid; mp 248-250 °C; IR (KBr) ν / cm-1 3492, 2974, 2934, 2866, 1462, 1386, 1004; 1H and 13C NMR data see Table 1; HRMS (ESI) (positive-ion mode) m/z, calcd. for C30H52NaO2+ [M + Na]+: 467.3859, found: 467.3856.

7β,26-Epoxyfriedelan-3a,7a-diol (3)

White solid; mp 145-147 °C; 1H and 13C NMR data see Table 1; HRMS (ESI) (positive-ion mode) m/z, calcd. for C30H50NaO3+ [M + Na]+: 481.3652, found: 481.3648.

3α-Hydroxyfriedelan-29-yl palmitate (4) and 3α-hydroxy­friedelan-29-yl stearate (5)

White solid; mp 126-130 °C; IR (KBr) ν / cm-1 3488, 2928, 2856, 1736, 1708, 1462, 1386, 1170, 408; 1H and 13C NMR data see Table 1; HRMS (APCI) (positive-ion mode) m/z, calcd. for C30H51O2+ ([M - 239]+ for 4 and [M - 267]+ for 5) 443.3884, found: 443.3886, m/z, calcd. for C16H31O+ 239.2369, found: 239.2368, m/z, calcd. C18H35O+ 267.2682, found: 267.2681; HRMS (APCI) (negative-ion mode): m/z, calcd. for C16H31O2- [M - terpenoid] -: 255.2330, found: 255.2328 (compound 4); calcd. for C18H31O2- [M - terpenoid]-: 283.2643, found: 283.2643 (compound 5).

3α-Hydroxyfriedelan-7-one (11)

White solid; mp 305-307 °C; IR (KBr) ν / cm-1 3482, 2942, 2868, 1704, 1456, 1388, 1036, 1006; 1H and 13C NMR data see Table 1; HRMS (ESI) (positive-ion mode) m/z, calcd. for C30H51O2+ [M + H]+: 443.3884; found: 443.3878.

Cytotoxicity evaluation

The cytotoxicity of the samples was evaluated against the human tumor cell lines THP-1 (human acute monocytic leukemia, ATCC-TIB-202), K562 (chronic myeloid leukemia, ATCC-CRL-3344), MDA-MB-231 (human cancer breast, ATCC-HTB-26) and TOV-21G (human cancer ovary, ATCC-HTB-26). The cytotoxicity of the samples against Wi-26VA4 cells (human lung embryonic fibroblast cells, ATCC- CCL-75) allowed to establish the selectivity index (SI). The cell viability was determined by colorimetric assay using 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide (MTT, Sigma-Aldrich, Saint Louis, USA). For the evaluation of cytotoxic activity, cells were placed in 96-well plates (1 × 106 cells per well) containing Roswell Park Memorial Institute 1640 (RPMI 1640) medium plus 10% fetal bovine serum (FBS). After the cell plating, the plates were incubated for 24 h at 37 ºC, in a 5% CO2 atmosphere, in a humid environment, for consecutive addition of the samples to be tested. The cytotoxicity assays were performed in four serial dilutions on the decimal scale from the stock solution (compounds and positive control in dimethyl sulfoxide-DMSO), using RPMI 1640 with 1% FBS supplementation. Each concentration was tested in triplicate and each test was also repeated in triplicate. Cytarabine (Blausiegel, Fortaleza, Brazil) (for THP-1), imatinib (Eurofarma, São Paulo, Brazil) (for K562) and etoposide (Sigma-Aldrich, Saint Louis, USA) (for MDA-MB-231 and TOV-21G) were used as positive controls. After 48 h of incubation, the MTT reagent (100 μL, 5 mg mL-1) was added to each well. After 3 h at 37 °C, 50 μL DMSO was added to each well to dissolve the formazan crystals. The colorimetric reading was measured at 550 nm using a SpectraMax Plus 384® microplate reader (Molecular Devices, San Jose, USA). The cytotoxicity was expressed by the sample concentration values that inhibit 50% of cell growth (IC50) in comparison to the cells cultivated in the absence of the sample (negative control). The selectivity index (SI) was calculated by the ratio of the IC50 obtained for normal cells (Wi-26VA4) to that obtained for the cancer cell lines.

Results and Discussion

Compound 1 was obtained as a white amorphous powder with molecular formula C30H50O2 as established by HR APCI-MS (m/z: 443.3886 [M + H]+, calcd. 443.3884). The IR spectrum displayed hydroxyl absorption at 3436 cm 11 Simmons, M. P.; Cappa, J. J.; Archer, R. H.; Ford, A. J.; Eichstedt, D.; Clevinger, C. C.; Mol. Phylogenet. Evol. 2008, 48, 745., C-O stretch at 1026 cm-1, C=C stretch at 1637 cm-1, and out-of-plane bending vibrations of =C-H at 994 and 907 cm-1. The 1H NMR spectrum showed seven singlets of methyl groups (δH 0.93; 0.94; 0.95; 0.99; 1.02; 1.11 and 1.18), three signals characteristic of a monosubstituted alkene at δH 5.60 (dd, 1H, J 17.4 and 10.8 Hz), δH 4.95 (d, 1H, J 10.8 Hz) and δH 4.89 (d, 1H J 17.4 Hz), and two signals at δH 5.09 (ddd, 1H, J 8.7, 6.1, 2.5 Hz) and δH 3.84 (dd, 1H, J 11.0, 5.3 Hz) (Table 1). The 13C spectral data were closely related to compound 13 (3,4-seco-friedelan-3,11β-olide),88 Sousa, G.; Duarte, L.; Alcântara, A.; Silva, G.; Vieira-Filho, S.; Silva, R.; Oliveira, D.; Takahashi, J.; Sousa, G. F.; Duarte, L. P.; Alcântara, A. F. C.; Silva, G. D. F.; Vieira-Filho, S. A.; Silva, R. R.; Oliveira, D. M.; Takahashi, J. A.; Molecules 2012, 17, 13439. suggesting a seco-friedelan with a possible hemiacetal ring involving C-3 (δC 95.8) and C-11 (δC 73.9). The HMBC spectrum demonstrated correlations of C-3 (δC 95.8) with the signals at δH 3.84 (H-11), δH 1.36 (H-1) and δH 1.52 (H-2), and of H-11 with the signals of C-10 (δC 62.3), C-12 (δC 39.4) and C-25 (δC 14.9), confirming the presence of both C-3 and C-11 in the same ring (Figure 2). The NOESY spectrum established the alpha position for H-3 and H-11 through the correlations of H-11 with H-8 (δH 1.35), H-10 (δH 1.0), H-12α (δH 1.50) and H-27 (δH 1.11), and of H-3 with H-2α (δH 2.08) and H-1 (δH 1.36), also implying in the beta position of the hydroxyl group linked to C-3. Furthermore, the 13C spectral data were also relatable to putranjivic acid,1414 Leong, Y.-W.; Harrison, L. J.; Phytochemistry 1999, 50, 849. especially regarding C-4 and C-23, implying a possible unsaturation between them. In the HSQC spectrum, H-23 (δH 4.95 and δH 4.89) correlated with C-23 (δC 111.1) and, in the COSY spectrum, with H-4 (δH 5.60). In the HMBC spectrum, the correlations of H-4 with C-5 (δC 42.0), C-10 and C-24 (δC 17.9) confirmed the presence of a double bond between C-4 and C-23 (Figures S1 to S10, Supplementary Information section Supplementary Information Supplementary information (IR, NMR spectra of compounds 1-16 and HR-ESI-MS/HR-APCI-MS analyses of compounds 1-5 and 11) (Figures S1-S89, Table S1) is available free of charge at http://jbcs.sbq.org.br as PDF file. ). The dataset allowed us to indicate compound 1 as the new friedelane triterpene of 3,4-seco-3,11β-epoxyfriedel-4(23)-en-3β-ol (1).

Figure 2
Correlations observed in COSY (1, (b) for compound 2.), HMBC () and NOESY (): (a) for compound

Compound 2 was isolated as a white amorphous powder with molecular formula C30H52O2 as established by HR ESI MS (m/z: 467.3856 [M + Na]+, calcd. 467.3859). The IR spectrum showed hydroxyl absorption at 3492 cm-1 and C-O stretch at 1004 cm-1. The 1H NMR spectrum displayed six singlets corresponding to seven methyl groups (δH 0.79, 3H; 0.87, 3H; 0.95, 3H; 0.99, 6H; 1.07, 3H; and 1.18, 3H), one doublet characteristic of a friedelane skeleton (H-23, δH 0.90, 3H, J 6.6 Hz), and signals attributed to carbinolic protons at δH 3.60 (dt, 3H, J 11.3 and 5.7 Hz) and δH 3.33 (td, H, J 10.7 and 4.6 Hz) (Table 1). The 13C NMR spectral data were related to friedelan-3β,11β-diol (16),1515 Sousa, G. F.; Ferreira, F. L.; Duarte, L. P.; Silva, G. D. F.; Messias, M. C.; Filho, S. V. A.; J. Chem. Res. 2012, 36, 203. with two signals at δC 71.9 and 76.9 from carbons bonded to hydroxyl groups. In the HMBC spectrum, the signal at δC 71.9 was attributed to C-3 according to the correlation with the doublet at δH 0.90 (H-23), which correlated with δC 53.4 (C-4) (Figure 2). The beta position of H-3 (δH 3.33) was confirmed by the NOESY spectrum correlations of this proton with the signals of H-24 (δH 0.79), H-23 (δH 0.90) and H-2β (δH 2.07). Thus, the hydroxyl group linked to C-3 was assigned at an alpha position, differently from the known compound 16. Moreover, the signal at δC 76.9 was attributed to C-11 due to the correlations of the signal at δH 3.60 (H 11) with δC 60.9 (C-10) and δC 13.3 (C-25) in the HMBC. In the NOESY spectrum, the alpha position of H-11 was confirmed with the with H-1α (δH 2.36), H-8 (δH 1.28), H-10 (δH 1.12), H-12α (δH 1.54), and H-27 (δH 1.07), implying the beta position of the C-11 hydroxyl group (Figures S11 to S23, Supplementary Information section Supplementary Information Supplementary information (IR, NMR spectra of compounds 1-16 and HR-ESI-MS/HR-APCI-MS analyses of compounds 1-5 and 11) (Figures S1-S89, Table S1) is available free of charge at http://jbcs.sbq.org.br as PDF file. ). The dataset allowed us to indicate compound 2 as the new friedelane triterpene friedelan-3α,11β-diol.

Compound 3 was isolated as a white solid with the molecular formula C30H50O3 established by HR-ESI-MS (m/z: 481.3648 [M + Na]+, calcd. 481.3652). The 1H NMR spectrum showed six singlets (Table 1) attributed to six methyl groups (δH 1.17, 1.04, 0.99, 0.94, 0.92 and 0.91), one doublet characteristic of H-23 of a friedelane skeleton (δH 1.02, 3H, J 6.5 Hz) and a signal attributed to a carbinolic proton at δH 3.56 (td, 1H, J 11.0, 6.6 Hz). The 13C NMR spectral data displayed three unshielded signals at δC 56.9, 74.6 and 77.4. As expected for friedelane compounds, the signal at δH 1.02 (H-23) correlated with C-3 (δC 74.6), C-4 (δC 53.3) and C-5 (δC 42.1) in the HMBC spectrum (Figure 3). Also, C-4 and C-5 correlated with H-6 (δH 1.74 and δH 2.01), and this proton further correlated with C-7 (δC 77.4), C-8 (δC 64.7) and C-10 (δC 58.1). Subsequent C-7 correlations with H-26 (δH 2.25 and δH 2.94) suggested a tetrahydropyran ring involving C-26 (δC 56.9) and C-7, further confirmed by the correlations of H-26 with C-8, C-14 (δC 42.7) and C-15 (δC 36.6). A similar ring was also observed in 24-hydroxy-3-oxohemiacetalfriedelane (rinol), isolated by Munvera et al.1616 Munvera, A. M.; Ouahouo, B. M. W.; Mkounga, P.; Mbekou, M. I. K.; Nuzhat, S.; Choudhary, M. I.; Nkengfack, A. E.; Nat. Prod. Res. 2020, 34, 2014. who also suggested this phenomenon as a probable result of a reaction between a ketone group at C-3 and a primary alcohol at C-24. The NOESY correlations of H-26 with H-25 (δH 1.17) and H-28 (δH 0.99) allowed to certify the beta position of the CH2O group, implying in the alpha position of the hydroxyl group at C-7 (Figures S24-S31, Supplementary Information section Supplementary Information Supplementary information (IR, NMR spectra of compounds 1-16 and HR-ESI-MS/HR-APCI-MS analyses of compounds 1-5 and 11) (Figures S1-S89, Table S1) is available free of charge at http://jbcs.sbq.org.br as PDF file. ). The dataset allowed us to indicate compound 3 as the new friedelane triterpene7β,26-epoxyfriedelan-3a,7a-diol.

Figure 3
Correlations observed in COSY (3, (b) for compounds 4 and 5.), HMBC () and NOESY (): (a) for compound

The mixture of compounds 4 and 5 was obtained as a white amorphous solid. The IR spectrum showed an alcohol O-H stretch band at 3488 cm-1 and an ester C=O stretch band at 1736 cm-1. The 1H NMR spectrum presented six singlets (δH 0.77, 0.81, 0.99, 1.00, 1.04 and 1.20) corresponding to six methyl groups and a characteristic methyl doublet signal related to H-23 of a friedelane skeleton at δH 0.89 (Table 1). This spectrum also showed two unshielded signals at δH 3.35 (td, 1H, J 10.4 and 4.8 Hz) and δH 3.74 (2H, m), probably belonging to protons bonded to an oxygenated carbon. The 13C NMR and distortionless enhancement by polarization transfer (DEPT-135) spectra presented one signal of an ester carbonyl at δc 174.3 and signals of methylene carbons between δc 29.2 and δc 30.0, possibly from a long chain ester. The 13C NMR data of the mixture were closely related to 3β,29-dihydroxyfriedelane.1717 Pereira, R. C. G.; Soares, D. C. F.; Oliveira, D. C. P.; de Sousa, G. F.; Vieira-Filho, S. A.; Mercadante-Simões, M. O.; Lula, I.; Silva-Cunha, A.; Duarte, L. P.; Magn. Reson. Chem. 2018, 56, 360. The HMBC correlations of the unshielded carbon at δC 72.2 with H-4 (δH1.05) and H-23 (δH 0.89), confirmed the hydroxyl group at C-3 (Figure 3). Additionally, H-3 (δH 3.35) showed correlations with H-2 (δH 1.25 and δH 2.07) and H-4 (δH 1.05) in the COSY spectrum. Furthermore, H-3 correlated with the signals of H-1 (δH 1.34), H-2β (δH 2.07), H-23 and H-24 (δH 0.77) in the NOESY spectrum, certifying the alpha position of the hydroxyl group linked to C-3 and the beta position of H-3. The ester group in C-29 was confirmed by correlations of the signal at δH 3.74 (H-29) with C-19 (δc 29.7), C-20 (δc 31.8), C-21 (δc 28.2), C-30 (δc 26.4), and C-1’ (δc 174.3) in the HMBC spectrum. The chemical shift assignments of the mixture of 4 and 5 were completely established through detailed analysis of HSQC, HMBC, COSY and NOESY spectra (Figures S32 to S42, Supplementary Information section Supplementary Information Supplementary information (IR, NMR spectra of compounds 1-16 and HR-ESI-MS/HR-APCI-MS analyses of compounds 1-5 and 11) (Figures S1-S89, Table S1) is available free of charge at http://jbcs.sbq.org.br as PDF file. ). In the positive mode, HR-APCI-MS spectrum of the mixture of 4 and 5 showed the fragment ion peak C30H51O2+ ([M-239]+ for 4 and [M - 267]+ for 5) resulting from the C29O-C1’ bond breaking. Also, this spectrum presented peaks referring to the acylium ions at m/z 239.2368 [C16H31O]+ (calcd. 239.2369) and at m/z 267.2681 [C18H35O]+ (calcd. 267.2682). In the negative mode, the HR-APCI-MS analysis displayed two ion peaks [M-triterpenoid]- at m/z 255.2328 and 283.2643, referring to palmitate and stearate ions, respectively. Therefore, the mixture was identified as 3α-hydroxyfriedelan-29-yl palmitate (4) and 3α-hydroxyfriedelan-29-yl stearate (5).

Compound 11 was identified by 1D/2D NMR spectra and HR-ESI-MS as 3α-hydroxyfriedelan-7-one (Figures S63 to S72, Supplementary Information section Supplementary Information Supplementary information (IR, NMR spectra of compounds 1-16 and HR-ESI-MS/HR-APCI-MS analyses of compounds 1-5 and 11) (Figures S1-S89, Table S1) is available free of charge at http://jbcs.sbq.org.br as PDF file. ). Even though the 1H NMR data for this compound is found in the literature,1818 Wittayalai, S.; Mahidol, C.; Prachyawarakorn, V.; Prawat, H.; Ruchirawat, S.; Phytochemistry 2014, 99, 121.

19 Sengupta, P.; Mukherjee, J.; Tetrahedron 1968, 24, 6259.
-2020 Sengupta, P.; Chakraborty, A. K.; Duffield, A. M.; Durham, L. J.; Djerassi, C.; Tetrahedron 1968, 24, 1205. this is the first report for its 13C NMR spectra data (Table 1). The other known compounds were identified as friedelan-3-one (6),2121 Mahato, S. B.; Kundu, A. P.; Phytochemistry 1994, 37, 1517. friedelan-3β-ol (7),2222 Salazar, G. C. M.; Silva, G. D. F.; Duarte, L. P.; Filho, S. A. V.; Lula, I. S.; Magn. Reson. Chem. 2000, 38, 977. friedelan-3α ol (8),2222 Salazar, G. C. M.; Silva, G. D. F.; Duarte, L. P.; Filho, S. A. V.; Lula, I. S.; Magn. Reson. Chem. 2000, 38, 977. friedelan-3,7-dione (9),2323 Patra, A.; Chaudhuri, S. K.; Magn. Reson. Chem. 1987, 25, 95. 3β-hydroxyfriedelan-7-one (10),2323 Patra, A.; Chaudhuri, S. K.; Magn. Reson. Chem. 1987, 25, 95. gutta-percha (12),2424 Tangpakdee, J.; Tanaka, Y.; Shiba, K.; Kawahara, S.; Sakurai, K.; Suzuki, Y.; Phytochemistry 1997, 45, 75. 3,4-seco-friedelan-3,11 olide (13),88 Sousa, G.; Duarte, L.; Alcântara, A.; Silva, G.; Vieira-Filho, S.; Silva, R.; Oliveira, D.; Takahashi, J.; Sousa, G. F.; Duarte, L. P.; Alcântara, A. F. C.; Silva, G. D. F.; Vieira-Filho, S. A.; Silva, R. R.; Oliveira, D. M.; Takahashi, J. A.; Molecules 2012, 17, 13439. β-sitosterol (14),2525 Kitajima, J.; Kimizuka, K.; Tanaka, Y.; Chem. Pharm. Bull. (Tokyo) 1998, 46, 1408. 11β-hydroxyfriedelan-3-one (15)2626 Chen, M.-X.; Wang, D.-Y.; Guo, J.; J. Chem. Res. 2010, 34, 114. and a mixture of friedelan-3α,11β-diol (2) and friedelan-3β,11β-diol (16)1515 Sousa, G. F.; Ferreira, F. L.; Duarte, L. P.; Silva, G. D. F.; Messias, M. C.; Filho, S. V. A.; J. Chem. Res. 2012, 36, 203. by comparison of their 1H and 13C NMR data with those reported in the literature. Compounds 9 and 10 were reported for the first time as metabolites of a Celastraceae species.

Cytotoxicity assay

The cytotoxic activity of the hexane extract (HE), the solid obtained during the hexane extraction (SHE) and compounds 1, 2, 6-11, 13, 15 were assessed against human leukemia (THP-1 and K562), ovarian and breast cancer cell lines (TOV-21G and MDA-MB-231, respectively). Compound 3 was not evaluated due to its small amount, and the cytotoxic activities of compounds 12 and 14 were previously reported in the literature.2424 Tangpakdee, J.; Tanaka, Y.; Shiba, K.; Kawahara, S.; Sakurai, K.; Suzuki, Y.; Phytochemistry 1997, 45, 75.,2525 Kitajima, J.; Kimizuka, K.; Tanaka, Y.; Chem. Pharm. Bull. (Tokyo) 1998, 46, 1408. Lung embryonic fibroblast cells (Wi-26VA4, ATCC-CCL-75) were adopted to establish the selectivity index (SI) of the analyzed compounds. Furthermore, etoposide was employed as the control with a further assessment using cytarabine for THP 1 and imatinib for K562 (Table 2). HE showed moderate cytotoxicity (IC50 < 43 ± 5 μg mL-1) against all tested cell lines, and SHE presented the best result against the THP-1 cell line (IC50 = 19 ± 1 μg mL-1, SI = 3).

Table 2
Cytotoxicity and selectivity of the hexane extract (HE), the solid obtained during the hexane extraction (SHE), and compounds 1, 2, 6-11, 13 and 15 against leukemia, ovarian and breast cancer cell lines

The results obtained for the leukemia cell lines (THP-1 and K562) demonstrated that all tested compounds showed IC50 values lower than 80 μg mL-1. Compounds 1, 2 and 15 displayed the best results (IC50 = 13 ± 1, 10.0 ± 0.9 and 14 ± 1 μg mL-1 and SI = 6, 5 and 4, respectively) against the THP-1 cell line. Also, compounds 2 and 15 presented good results against the K562 cell line (IC50 = 11 ± 1 and 16 ± 2 μg mL-1 and SI = 4 and 3, respectively). Concerning the ovarian and breast cancer cell lines, compound 2 was the most active and selective against TOV-21G and MDA MB-23 cell lines (IC50 = 33 ± 3 and 13 ± 2 μg mL-1, and SI = 2 and 4, respectively).

Indeed, observing the structures of compounds 2 and 15, which displayed the best results against both THP-1 and K562, suggested that a hydroxyl group in C-11 may be relevant to the antileukemic activity. On the other hand, only compound 2, which displayed a hydroxyl group at C-3, demonstrated good results against the ovarian and breast cancer cell lines. The results corroborated with the literature on triterpenes inhibiting cancer cell proliferation.2727 Dash, S. K.; Giri, B.; J. Exp. Med. 2018, 1, 1.

28 Shanmugam, M. K.; Dai, X.; Kumar, A. P.; Tan, B. K. H.; Sethi, G.; Bishayee, A.; Cancer Lett. 2014, 346, 206.
-2929 Morales, S. A. T.; de Aguilar, M. G.; Pereira, R. C. G.; Duarte, L. P.; Sousa, G. F.; de Oliveira, D. M.; Evangelista, F. C. G.; Sabino, A. P.; Viana, R. O.; Alves, V. S.; Vieira-Filho, S. A.; Quim. Nova 2020, 43, 1066. Pereira et al.3030 Pereira, R. C. G.; Evangelista, F. C. G.; Santos Jr., V. S.; Sabino, A. P.; Maltarollo, V.; Freitas, R. P.; Duarte, L. P.; Chem. Biodiversity 2020, 17, e2000773. evaluated a series of pentacyclic triterpenes against leukemic cell lines (THP-1 and K562) and the best results were found for friedelan-3β-ol (7) and α-amyrin. Employing chemometric analysis, the authors observed that the activity could be related to the presence of a hydroxyl group at C-3 and its stereoelectronic interactions with molecular targets.3030 Pereira, R. C. G.; Evangelista, F. C. G.; Santos Jr., V. S.; Sabino, A. P.; Maltarollo, V.; Freitas, R. P.; Duarte, L. P.; Chem. Biodiversity 2020, 17, e2000773. Thus, further structure-activity relationships studies should be conducted for a deeper understanding.

Conclusions

The phytochemical study of the hexane extract from Maytenus quadrangulata leaves led to the isolation of 14 triterpenes, one steroid and the polymer gutta-percha. The 1D/2D NMR spectral data of 3,4-seco-3,11β epoxyfriedel-4(23)-en-3β-ol (1), friedelan-3α,11β diol (2), 7β,26-epoxyfriedelan-3a,7a-diol (3) and a mixture of 3α-hydroxyfriedelan-29-yl palmitate (4) and 3α-hydroxyfriedelan-29-yl stearate (5) are herein described for the first time. Compound 2 demonstrated the best results among all tested samples, showing promising cytotoxicity and selectivity against THP-1, K562, TOV-21G and MDA MB-231 cell lines.

Supplementary Information

Supplementary information (IR, NMR spectra of compounds 1-16 and HR-ESI-MS/HR-APCI-MS analyses of compounds 1-5 and 11) (Figures S1-S89, Table S1) is available free of charge at http://jbcs.sbq.org.br as PDF file.

Acknowledgments

The authors thank Fundação de Amparo à Pesquisa do Estado de Minas Gerais, Coordenação de Aperfeiçoamento de Pessoal de Nível Superior-Brasil (CAPES), Pró-Reitoria de Pesquisa da Universidade Federal de Minas Gerais for the financial support, Centro Regional para o Desenvolvimento Tecnológico e Inovação (UFG) and Centro de Pesquisa René Rachou (FIOCRUZ) for the HRMS analysis, botanist Dr Maria Olívia Mercadante-Simões (UNIMONTES) for the first collection and identification of the plant and MSc Salomão B. V. Rodrigues (UFMG) for the English revision and critical reading of the manuscript.

References

  • 1
    Simmons, M. P.; Cappa, J. J.; Archer, R. H.; Ford, A. J.; Eichstedt, D.; Clevinger, C. C.; Mol. Phylogenet. Evol. 2008, 48, 745.
  • 2
    Christenhusz, M. J. M.; Byng, J. W.; Phytotaxa 2016, 261, 201.
  • 3
    Mokoka, T. A.; McGaw, L. J.; Mdee, L. K.; Bagla, V. P.; Iwalewa, E. O.; Eloff, J. N.; BMC Complementary Altern. Med. 2013, 13, 111.
  • 4
    Zhang, L.; Ji, M.-Y.; Bin, Q.; Li, Q.-Y.; Zhang, K.-Y.; Liu, J.-C.; Dang, L.-S.; Li, M.-H.; Med. Chem. Res. 2020, 29, 575.
  • 5
    Cardoso, D. B. O. S.; de Queiroz, L. P.; J. Bot. Res. Inst. Texas 2008, 2, 551.
  • 6
    Menino, G. C. O.; dos Santos, R. M.; Apgaua, D. M. G.; Pires, G. G.; Pereira, D. G. S.; Fontes, M. A. L.; Almeida, H. S.; CERNE 2015, 21, 277.
  • 7
    Celastraceae in Flora do Brasil 2020, Jardim Botânico do Rio de Janeiro, http://floradobrasil.jbrj.gov.br/reflora/floradobrasil/FB16751, accessed in March 2022.
    » http://floradobrasil.jbrj.gov.br/reflora/floradobrasil/FB16751
  • 8
    Sousa, G.; Duarte, L.; Alcântara, A.; Silva, G.; Vieira-Filho, S.; Silva, R.; Oliveira, D.; Takahashi, J.; Sousa, G. F.; Duarte, L. P.; Alcântara, A. F. C.; Silva, G. D. F.; Vieira-Filho, S. A.; Silva, R. R.; Oliveira, D. M.; Takahashi, J. A.; Molecules 2012, 17, 13439.
  • 9
    Xu, C.; Wang, B.; Pu, Y.; Tao, J.; Zhang, T.; J. Sep. Sci. 2017, 41, 6.
  • 10
    Antonisamy, P.; Duraipandiyan, V.; Aravinthan, A.; Al-Dhabi, N. A.; Ignacimuthu, S.; Choi, K. C.; Kim, J.-H.; Eur. J. Pharmacol. 2015, 750, 167.
  • 11
    Antonisamy, P.; Duraipandiyan, V.; Ignacimuthu, S.; J. Pharm. Pharmacol. 2011, 63, 1070.
  • 12
    Sunil, C.; Duraipandiyan, V.; Ignacimuthu, S.; Al-Dhabi, N. A.; Food Chem. 2013, 139, 860.
  • 13
    Aguilar, M. G.; Sousa, G. F.; Evangelista, F. C. G.; Sabino, A. P.; Vieira Filho, S. A.; Duarte, L. P.; Nat. Prod. Res. 2020, 34, 810.
  • 14
    Leong, Y.-W.; Harrison, L. J.; Phytochemistry 1999, 50, 849.
  • 15
    Sousa, G. F.; Ferreira, F. L.; Duarte, L. P.; Silva, G. D. F.; Messias, M. C.; Filho, S. V. A.; J. Chem. Res. 2012, 36, 203.
  • 16
    Munvera, A. M.; Ouahouo, B. M. W.; Mkounga, P.; Mbekou, M. I. K.; Nuzhat, S.; Choudhary, M. I.; Nkengfack, A. E.; Nat. Prod. Res. 2020, 34, 2014.
  • 17
    Pereira, R. C. G.; Soares, D. C. F.; Oliveira, D. C. P.; de Sousa, G. F.; Vieira-Filho, S. A.; Mercadante-Simões, M. O.; Lula, I.; Silva-Cunha, A.; Duarte, L. P.; Magn. Reson. Chem. 2018, 56, 360.
  • 18
    Wittayalai, S.; Mahidol, C.; Prachyawarakorn, V.; Prawat, H.; Ruchirawat, S.; Phytochemistry 2014, 99, 121.
  • 19
    Sengupta, P.; Mukherjee, J.; Tetrahedron 1968, 24, 6259.
  • 20
    Sengupta, P.; Chakraborty, A. K.; Duffield, A. M.; Durham, L. J.; Djerassi, C.; Tetrahedron 1968, 24, 1205.
  • 21
    Mahato, S. B.; Kundu, A. P.; Phytochemistry 1994, 37, 1517.
  • 22
    Salazar, G. C. M.; Silva, G. D. F.; Duarte, L. P.; Filho, S. A. V.; Lula, I. S.; Magn. Reson. Chem. 2000, 38, 977.
  • 23
    Patra, A.; Chaudhuri, S. K.; Magn. Reson. Chem. 1987, 25, 95.
  • 24
    Tangpakdee, J.; Tanaka, Y.; Shiba, K.; Kawahara, S.; Sakurai, K.; Suzuki, Y.; Phytochemistry 1997, 45, 75.
  • 25
    Kitajima, J.; Kimizuka, K.; Tanaka, Y.; Chem. Pharm. Bull. (Tokyo) 1998, 46, 1408.
  • 26
    Chen, M.-X.; Wang, D.-Y.; Guo, J.; J. Chem. Res. 2010, 34, 114.
  • 27
    Dash, S. K.; Giri, B.; J. Exp. Med. 2018, 1, 1.
  • 28
    Shanmugam, M. K.; Dai, X.; Kumar, A. P.; Tan, B. K. H.; Sethi, G.; Bishayee, A.; Cancer Lett. 2014, 346, 206.
  • 29
    Morales, S. A. T.; de Aguilar, M. G.; Pereira, R. C. G.; Duarte, L. P.; Sousa, G. F.; de Oliveira, D. M.; Evangelista, F. C. G.; Sabino, A. P.; Viana, R. O.; Alves, V. S.; Vieira-Filho, S. A.; Quim. Nova 2020, 43, 1066.
  • 30
    Pereira, R. C. G.; Evangelista, F. C. G.; Santos Jr., V. S.; Sabino, A. P.; Maltarollo, V.; Freitas, R. P.; Duarte, L. P.; Chem. Biodiversity 2020, 17, e2000773.

Edited by

Editor handled this article: Paulo Cezar Vieira

Publication Dates

  • Publication in this collection
    24 Oct 2022
  • Date of issue
    Nov 2022

History

  • Received
    08 Dec 2021
  • Accepted
    01 Apr 2022
Sociedade Brasileira de Química Instituto de Química - UNICAMP, Caixa Postal 6154, 13083-970 Campinas SP - Brazil, Tel./FAX.: +55 19 3521-3151 - São Paulo - SP - Brazil
E-mail: office@jbcs.sbq.org.br