Skip to main content

Study on the absolute configuration and biological activity of rotenoids from the leaves and twigs of Millettia pyrrhocarpa Mattapha, Forest & Hawkins, sp. Nov

Abstract

Background

M. pyrrhocarpa is a new plant in the Fabaceae: Faboideae family that is found in Thailand. A literature search revealed that the Milletia genus is rich in bioactive compounds possessing a wide range of biological activities. In this study, we aimed to isolate novel bioactive compounds and to study their bioactivities.

Methods

The hexane, ethyl acetate, and methanol extracts from the leaves and twigs of M. pyrrhocarpa were isolated and purified using chromatography techniques. These extracts and pure compounds were tested in vitro for their inhibitory activities against nine strains of bacteria, as well as their anti-HIV-1 virus activity and cytotoxicity against eight cancer cell lines.

Results

Three rotenoids, named 6aS, 12aS, 12S-elliptinol (1), 6aS, 12aS, 12S-munduserol (2), dehydromunduserone (3), and crude extracts were evaluated for antibacterial, anti-HIV, and cytotoxic activities. It was found that compounds 1–3 inhibited the growth of nine strains of bacteria, and the best MIC/MBC values were obtained at 3/ > 3 mg/mL. The hexane extract showed anti-HIV-1 RT with the highest %inhibition at 81.27 at 200 mg/mL, while 6aS, 12aS, 12S-elliptinol (1) reduced syncytium formation in 1A2 cells with a maximum EC50 value of 4.48 μM. Furthermore, 6aS, 12aS, 12S-elliptinol (1) showed cytotoxicity against A549 and Hep G2 cells with maximum ED50 values of 2.27 and 3.94 μg/mL.

Conclusion

This study led to the isolation of constituents with potential for medicinal application, providing compounds (1–3) as lead compounds against nine strains of bacteria. The hexane extract showed the highest %inhibition of HIV-1 virus, Compound 1 showed the best EC50 in reducing syncytium formation in 1A2 cells, and it also showed the best ED50 against human lung adenocarcinoma (A549) and human hepatocellular carcinoma (Hep G2). The isolated compounds from M. pyrrhocarpa offered significant potential for future medicinal application studies.

Peer Review reports

Background

Millettia pyrrhocarpa Mattapha, Forest & Hawkins, sp. Nov. belongs to the family Fabaceae: Faboideae and is found in Nakorn Nayok province, Thailand. M. pyrrhocarpa is a woody climber plant that is known in Thai as “Nang Rong” [1]. Previous phytochemical analyses of the Milletia genus revealed that various secondary metabolites were present, such as alkaloids [2], coumarins [3], flavonoids [4], isoflavonoids [5], phenols [6], phytosterols [2], rotenoids [710], and triglycerides [11]. Among these, rotenoids are the most common in this genus [12], as observed in the leaves of Millettia oblate ssp. teitensis [13], Millettia brandisiana KURZ [14] and in the root bark of Millettia usaramensis [15] and Millettia speciosa [16]. Pharmacological investigation of the Millettia genus revealed that the crude extract and isolated compounds showed antimicrobial [2, 17, 18], antioxidant [2, 18], antiplasmodial [7, 10], immunomodulatory [19], anti-cholinesterase [20, 21], anthelmintic [22], anti-inflammatory [5, 23], antidiabetic [24], cytotoxic [6, 16, 20], and anticancer [6, 20, 25] activities. Since the phytochemical study and biological activity of M. pyyrhocarpa has not been reported, it is of great interest to study the secondary metabolites in the leaves and twigs of M. pyyrhocarpa and their antibacterial, anti-HIV1-RT, and cytotoxic activities. The preliminary evaluation of the bioactivity of the leaves and twigs of M. pyyrhocarpa might lead to the discovery of important substances that exhibit promising biological activity and potential to be developed for medicinal applications in the future.

Material

Plant material

Leaves and twigs of Millettia pyrrhocarpa were collected from Nakhon Nayok Province, Mueang district, Hin Tang subdistrict, Khao Yai National Park, Nang Rong waterfall in Thailand and identified by Dr. Narong Nantasean. Voucher specimens of M. pyrrhocarpa (M. pyrrhocarpa Mattapan, & BKF staff 1139B holotype BKF) were deposited at the Department of Chemistry, Faculty of Science and Center of Innovation in Chemistry, Mahidol University, Rama VI Road, Bangkok 10400, Thailand.

Cells line materials KKU-M213 (human cholangiocarcinoma) cells were kindly provided by Dr. Banchob Sripa from Liver Fluke and Cholangiocarcinoma Research Center, Department of Pathology, Faculty of Medicine, Khon Kean University. MMNK1 (human cholangiocyte) cells were obtained from the Japanese Collection of Research Bioresources Cell Bank (JCRB, Osaka, Japan). FaDu (human hypopharyngeal carcinoma), HT29 (human colorectal adenocarcinoma), MDA-MB-231 (human mammary gland/breast adenocarcinoma), SH-SY5Y (human neuroblastoma), A549 (human lung carcinoma), and HepG2 (human hepatocellular carcinoma) cell lines were obtained from the American Type Culture Collection (ATCC, Manassas, VA, USA).

Methods

General experimental procedures

Column chromatography (CC) was carried out on silica gel 60 H from E. Merck. 70–230 mesh ASTM, cat. No. 7734 and No.7736. Thin-layer chromatography (TLC) separations were carried out on silica gel 60 PF254 on aluminium sheets, and the isolated compounds were identified under ultraviolet light. Infrared spectra (IR) were recorded as KBr pellets using a Shimadzu 8900 FT-IR spectrophotometer. Melting points were recorded on a Büchi 322 micro melting point apparatus and are uncorrected. Mass spectra were recorded on a Thermo Finnigan Polaris Q mass spectrometer at 70 eV (probe), and EIMS was measured by a Brüker Esquire apparatus. X-ray absorption spectroscopy was carried out on a Bruker D8 QUEST CMOS PHOTON II. 1H (500 MHz), 13C (125 MHz), and 2D NMR spectra were recorded on a Bruker AV-500 spectrometer in deuterated chloroform (CDCl3) or deuterated methanol (CD3OD) solutions, and TMS was used as the internal standard.

Extraction and isolation

Air-dried powders of the leaves and twigs of M. pyrrhocarpa (2.8 kg) were extracted successively with hexane, EtOAc, and MeOH. Twelve litres of solvent were used in the extraction at room temperature for three days in triplicate. The solvents were filtered and evaporated under reduced pressure to give crude hexane (41.0 g), EtOAc (118.0 g), and MeOH (216.0 g) extracts. The crude hexane extract was fractionated using silica gel column chromatography (CC) using a gradient system of hexane–EtOAc and EtOAc–MeOH to generate ten fractions (A1-A10). Fraction A5 (13.6 g) was fractionated on a silica gel column and eluted with a gradient system of hexane–EtOAc and EtOAc–MeOH to give three subfractions (B1-B3). Subfraction B2 (5.1 g) was fractionated on a silica gel column and eluted with a gradient system of hexane–ethyl acetate and EtOAc–MeOH to give four subfractions, C1-C4. Subfraction C1 (0.06 g) was recrystallized with ethanol:ethyl acetate (2:1) to give compound (1) (70 mg) as a white solid. Subfraction C3 (0.41 g) was recrystallized with ethanol:ethyl acetate (2:1) to afford compound (3) (40 mg) as a white solid. The crude EtOAc extract (118.0 g) was fractionated on a silica gel column and eluted with a gradient system of hexane–EtOAc and EtOAc–MeOH to afford nine fractions (D1-D9). Fraction D2 (23.3 g) was fractionated on a silica gel column and eluted with a gradient system of hexane–ethyl acetate and EtOAc–MeOH to produce seven subfractions (E1-E7). Subfraction E4 (3.0 g) was recrystallized with ethanol:ethyl acetate (2:1) to generate compound (1) (86 mg) as a white solid. The crude MeOH extract (216.1 g) was fractionated on a silica gel column and eluted with a gradient system of hexane–EtOAc and EtOAc–MeOH to generate nine fractions (G1-G9). Fraction G3 (1.3 g) was recrystallized with ethanol to afford compound (2) (36 mg) as a white solid.

X-ray diffraction

The colourless plate of compound (2) was suitable for single-crystal X-ray diffraction with a size of 0.32 × 0.28 × 0.04 mm. The unit cell parameters and intensity data were recorded on a Bruker D8 QUEST CMOS PHOTON II diffractometer equipped with a graphite-monochromator Mo-Kα (λ = 071,073 Å) radiation at 296(2) K. Data reduction was performed using SAINT, and the SADABS-2016/2 scaling algorithm [26] was used for absorption correction. The structure was solved with the ShelXT structure solution program using combined Patterson and dual-space recycling methods [27]. The structure was refined by least squares using ShelXL [28]. All non-H atoms were refined anisotropically. The hydrogen atoms of solvent molecules were positioned geometrically with C—H = 0.93–0.98 Å and refined using a rigid model with fixed displacement parameters Uiso (H) = 1.5Ueq (C) for methyl groups and 1.2Ueq (C) for the other groups. The O–H hydrogen atoms were located on difference Fourier maps but refined with O–H = 0.82 ± 0.02 Å with Uiso (H) = 1.5Ueq (O). C19H20O6, FW = 344.35, orthorhombic space Group P21P21P21, unit cell dimensions a = 6.40000(10) Å, b = 9.9684(2) Å, c = 25.9394(6) Å, V = 1654.88(6) Å3, Z = 4, dcalcd = 1.382 g/cm3, µ = 0.103 mm−1, F(000) = 728. The 46671 measurements yielded 6568 independent reflections after equivalent data were averaged. The final refinement gave R1 = 0.0393 and wR2 = 0.0896 [I > 2σ(I)]. The crystallographic data of the compound have been deposited in the Crystallographic Open Database (COD) number 3000415. The molecular graphic was illustrated by ORTEP [29].

Antibacterial activity

Bacterial strains

The study on in vitro antibacterial activity was carried out against nine strains (S. aureus ATCC 25923 DMST 8840, E. aerogenes ATCC13048 DMST 8841, E. coli O157: H7 DMST 12743, E. coli Enterotoxigenic, ETEC DMST 30543, E. coli Enteropathogenic, EPEC DMST 30546, S. typhimurium ATCC 13311 DMST 562, S. flexneri DMST 4423, P. mirabilis DMST 8212, and V. cholera nonO1/nonO139 DMST 2873). The minimum inhibitory concentration (MIC) and minimum bactericidal concentration (MBC) of the crude extracts and isolated compounds were evaluated according to standard methods described in previous literature [30].

Anti-HIV1-RT and cell-based assay for anti-HIV-1

Anti-HIV1-RT assays of the crude extracts and isolated compounds were evaluated according to the standard methods described in previous literature [3133]. Cell-based assays for anti-HIV-1 RT of the crude extracts and isolated compounds were evaluated according to the standard methods described in previous literature [34, 35]

Cytotoxicity assay

Cytotoxicity assays of the crude extracts and isolated compounds were evaluated according to the standard methods described in previous literature [36, 37].

Results

Isolation and purification

The crude hexane, ethyl acetate, and methanol extracts of the leaves and twigs of M. pyrrhocarpa were subjected to repeated chromatography over silica gel 60 and silica gel 60 PF254 to yield three pure compounds, (1–3). Compounds (1–3) were identified as elliptinol (1) [38, 39], munduserol (2) [40], and dehydromunduserone (3) [40] by comparison of their spectral data with those in the literature. The 1H and 13C-NMR data of. The 1H and 13C-NMR data of (1), (2), and (3) are shown in Table 1.

Table 1 1H and 13C data for compounds (1–3) (500 MHz, 125 MHz, CDCl3 or CD3OD, δ in ppm)

6aS, 12aS, 12S-Elliptinol (1): white solid (EtOH:EtOAc); mp = 230 − 231 °C;  \({\left[\alpha \right]}_{589}^{26.9}\) : + 173.36 (c 0.38, CHCl3); UV (MeOH) λmax (log ε) 235 (1.37), 281 (1.24), 341 (1.09), 447 (1.35), 494 (0.93), 671 (0.91) nm; IR (KBr) υmax 3470, 3102, 2969, 2938, 1620, 1586, 1502, 1273, 1253, 1190, 1142, 1119, 1092, 1041, 989 cm−1; for 1H and 13C NMR spectroscopic data, see Table 1. EIMS m/z 340.31 (base peak), 226.22, 81.05.

6aS, 12aS, 12S-Munduserol (2): white solid (EtOH:EtOAc); mp = 240 − 241 °C; \({\left[\alpha \right]}_{589}^{\text{26.}{2}}\) : + 160.03 (c 0.14, MeOH); UV (MeOH) λmax (log ε) 231 (1.73), 450 (1.33), 507 (1.23), 680 (1.19) nm; IR (KBr) υmax 3460, 2918, 2851, 1655, 1637, 1618, 1508, 1292, 1157, 1045 cm−1; for 1H and 13C NMR spectroscopic data, see Table 1. EIMS m/z 345.31 (M + H)+, 192.17 (base peak).

Dehydromunduserone (3): white solid (MeOH); mp = 202–204 °C [lit. 41, mp = 210 − 211 °C]; UV (MeOH) λmax (log ε) 231 (1.59), 275 (1.46), 306 (1.36), 356 (1.12), 457 (1.01), 507 (0.92), 677 (0.90) nm; IR (KBr) υmax 3441, 3098, 2999, 2955, 2837, 1761, 1726, 1634, 1601, 1566, 1508, 1452, 1441, 1408, 1290, 1246, 1199, 1163, 1105, 1049, 795 cm−1; for 1H and 13C NMR spectroscopic data, see Table 1. EIMS m/z 354.37, 192.29 (base peak), 179.31.

Antibacterial activity

The three crude extracts and isolated compounds (1–3) were found to inhibit the growth of V. cholerae with MIC/MBC values of 12.5/25 mg/mL, while the EtOAc extract inhibited the growth of E. coli (ETEC), E. coli (EPEC), and S. flexneri with MIC/MBC values of 6.25/6.25 mg/mL, and the MeOH extract was found to inhibit the growth of E. coli (ETEC) with MIC/MBC values of 6.25/6.25 mg/mL. The isolated compounds (1–3) were found to inhibit the growth of nine bacterial strains with MIC/MBC values of 3 < p3 mg/mL with chloramphenicol as the positive control. These results are shown in Figs. 1 and 2 and were consistent with previous reports [2, 17, 18].

Fig. 1
figure 1

Determination of MIC for crude extracts and pure compounds from leaves and twigs of M. Pyrrhocarpa

Fig. 2
figure 2

Determination of MBC for crude extracts and pure compounds from leaves and twigs of M. Pyrrhocarpa

Anti-HIV-1 RT activity

The hexane and EtOAc extracts showed the highest %inhibition of HIV-1 RT, at values of 81.27 and 66.97 at 200 µg/mL, while the MeOH extract was inactive. Compounds (1) and (2) reduced syncytium formation in 1A2 cells at EC50 values of 4.48 and 4.99 μM, respectively. AZT was used as a positive control, as shown in Figs. 3 and 4.

Fig. 3
figure 3

Anti-HIV-1 RT of crude extracts and pure compounds from leaves and twigs of M. Pyrrhocarpa. aAnti-HIV-1 RT activity express as %inhibition at 200 µg/mL: very active (VA) =  > 70% inhibition, moderately active (MA) = 50% to 69% inhibition, weakly active (WA) = 30% to 50% inhibition and inactive (I) =  < 30% inhibition; For determination of IC50 in the HIV-1 RT assay, the coefficients of determination, R2, were 0.98–0.99 in all assays for 50% end point

Fig. 4
figure 4

Anti-syncytium MC99 + 1A2 of crude extracts and pure compounds from leaves and twigs of M. Pyrrhocarpa. bAnti-syncytium (MC99 + 1A2) EC50 = dose of compound that reduced 50% syncytium formation by ΔTat/RevMC99 virus in 1A2 cells. AZT, averaged from three experiments, EC50 3.95 × 10− 3 μM;.cTI, Therapeutic Index: IC50/EC50

Cytotoxicity activity

The hexane extract showed cytotoxicity against KKU-M213, FaDu, HT-29, A549, SH-SY5Y, MNN-K1, and Hep G2 cells at ED50 values of 3.42, 6.20, 3.37, 4.45, 7.21, 12.63, and 1.23 µg/mL, respectively, while the ethyl acetate extract showed cytotoxicity against KKU-M213, HT-29, A549, SH-SY5Y, MNN-K1, and Hep G2 cells at ED50 values of 6.20, 14.99, 8.84, 11.79, 11.00 and 3.31 µg/mL, respectively. 6aS, 12aS, 12S-elliptinol (1) showed cytotoxicity against A549 and HepG2 at ED50 values of 2.27 and 3.97 µg/mL. Ellipticine [41, 42] was used as a positive control, as shown in Fig. 5.

Fig. 5
figure 5

Cytotoxicity study of crude extracts and isolated compounds from leaves and twigs of M. Pyrrhocarpa. dCytotoxic assay: ED50 less than 20 µg/mL were considered active for extracts and ED50 less than 4 µg/mL were considered active for pure compounds. Cancer cell lines: KKU-M213 (Human intrahepatic cholangiocarcinoma) FaDu (Human squamous cell carcinoma) HT-29 (Human colon adenocarcinoma) MDA-MB-231 (Human mammary gland/breast adenocarcinoma) A 549 (Human lung adenocarcinoma) SH-SY5Y (Human neuroblastoma) MNN-K1 (highly differentiated immortalized human cholangiocyte cell line) Hep G2 (Human hepatocellular carcinoma)

Discussion

Structure elucidation

Although the spectral data of compounds (1) and (2) were consistent with previous reports [3840], it was noted that the coupling constants and proton orientation of compounds (1) and (2) at positions 6a, 12a, and 12 of MOM-protected munduserol [43], elliptinol [38, 39], 6a, 12a-cis-12, 12a-cis-12-acetoxy-6,6a,12,12a-tetrahydrorotoxen [44], compounds (1), and (2) are quite different, and the structures and 1H-NMR data of the four compounds are shown in Fig. 6 and Table 2.

Fig. 6
figure 6

Structures of MOM-protected munduserol [43], elliptinol [38, 39], and 6a, 12a-cis-12, 12a-cis-12-acetoxy-6,6a, 12, 12a-tetrahydrorotoxen [44]

Table 2 Comparison of chemical shift and coupling constant values of MOM-protected munduserol [43], elliptinol [38, 39], 6a, 12a-cis-12, 12a-cis-12-acetoxy-6,6a,12,12a-tetrahydrorotoxen [44], compound (1) and (2)

From Table 2, the coupling constant between H12a and H12 was found to be 10.2 Hz in compound (1) and 10.0 Hz in compound (2), which suggested that H12a and H12 in compound (1) and compound (2) were trans to each other. This conclusion was contradictory to the findings in previous reports [38, 39, 43, 44]. To settle this discrepancy, compound (2) was subjected to X-ray analysis, which confirmed that H6a, H12a and H12 were in fact H6aβ, H12aα, and H12β, respectively. The absolute configuration of compound (2) was therefore 6aS, 12aS, 12S. Comparison of the 1H-NMR data of compound (1) with those of 6aS, 12aS, 12S-munduserol (2) led to the conclusion that compound (1) was 6aS, 12aS, 12S-elliptinol (1). In this work, 6aS, 12aS, 12S-elliptinol (1) and 6aS, 12aS, 12S-munduserol (2) from the leaves and twigs of M. pyrrhocarpa afforded useful spectroscopic and XRD data, making it possible to assign the protons at the 6a, 12a, and 12 positions. The key COSY ( ) and HMBC ( ) correlations of compounds (1–3) are shown in Fig. 7, and the X-ray crystal structure of compound (2) is shown in Fig. 8.

Fig. 7
figure 7

Key COSY ( ) and HMBC ( ) correlations of 6aS, 12aS, 12S-elliptinol (1), 6aS, 12aS, 12S-munduserol (2) and dehydromunduserone (3)

Fig. 8
figure 8

X-ray crystal structure of 6aS, 12aS, 12S-munduserol (2)

Bioactivities

Regarding antibacterial activity, the crude EtOAc extract from M. pyrrhocarpa showed good antibacterial activity against E. Coli and S. flexneri at MIC and MBC values ​​of 6.25 mg/mL. Compounds (1–3) inhibited 9 strains of bacterial cells at MIC and MBC values of 3 and > 3 mg/mL, which was more prominent than other Milletia genera tested in previous reports [45, 46]. In anti-HIV activity, compounds (1) and (2) from M. pyrrhocarpa showed anti-syncytium activity with better IC50 and EC50 values than that of the pure compound found in Ventilago harmandiana [47]. In in vitro cytotoxic activity, crude hexane, ethyl acetate, and compound (1) from M. pyrrhocarpa showed cytotoxicity against A549 (human lung adenocarcinoma) cells with better ED50 values at 4.45, 8.84, and 2.27 µg/mL than those of the crude extract and pure compounds from Garcinia speciosa Wall [48]. In addition, compared to phenylacetylshikonin analogues, compound (1) showed comparable and better cytotoxicity to A549 cells [49]. Extracts and purified compounds (1–3) from M. pyrrhocarpa showed bioactivity, especially compound (1), which showed potent anticancer activity. This work does possess inherent limitations. The scarcity of these compounds makes it impossible to test other biological activities, such as anti-inflammatory and antidiabetic activities, and perform in vivo studies. Further molecular modelling and computational studies, such as molecular docking and molecular dynamics techniques, are desirable. Through these studies, the interaction between the tested compounds and cells could be predicted to obtain suitable structures so that appropriate syntheses can be carried out.

Conclusion

This is the first phytochemical investigation of the leaves and twigs of M. pyrrhocarpa and the first study to examine their biological activity. Interestingly, previous reports did not establish the absolute configuration of rotenol (1) and (2) [38,39,40]. This report established the complete structure, spectral data of compounds (1) and (2), and X-ray data of (2). Furthermore, bioassays such as antibacterial, anti-HIV virus, and cytotoxicity on rotenol (1), (2), and rotenone (3) provided significant data that could be used for further study.

Availability of data and materials

The datasets generated and analysed during the current study are available in the corresponding author and the Crystallographic Open Database (COD) number 3000415 repository, [http://www.crystallography.net/cod/result.php].

References

  1. Mattapha S, Forest F, Hawkins J, Suddee S, Tetsana N, Chantaranothai P. Three new species, lectotypifications and synonymisations in Millettia (Fabaceae: Faboideae) for Thailand. Thai For Bull (Bot). 2019;47(2):171–3. https://doi.org/10.20531/tfb.2019.47.2.07.

    Article  Google Scholar 

  2. Havyarimana L, Ndendoung ST, Tamokou JD, Atchadé AT, Tanyi JM. Chemical constituents of Millettia barteri and their antimicrobial and antioxidant activities. Pharm Biol. 2012;50(2):141–6. https://doi.org/10.3109/13880209.2011.579618.

    Article  CAS  PubMed  Google Scholar 

  3. Rajemiarimiraho M, Banzouzi JT, Rakotonandrasana SR, Chalardd P, Benoit-Vicale F, Rasoanaivog LH, Raharisololalaog A, Randrianja R. Pyranocoumarin and triterpene from Millettia Richardiana. Nat Prod Commun. 2013;8(8):1099.

    CAS  PubMed  Google Scholar 

  4. Rahman TU, Arfan M, Mahmood T, Liaqat W, Gilani MA, Uddin G, Ludwig R, Zaman K, Choudhary MI, Khattak KF, Ayub K. Isolation, spectroscopic and density functional theory studies of 7-(4-methoxyphenyl)-9H-furo[2,3-f]chromen-9-one: a new flavonoid from the bark of Millettia ovalifolia. Spectrochim Acta-A: Mol Biomol Spectrosc. 2015;146:24–32. https://doi.org/10.1016/j.saa.2015.03.061.

    Article  CAS  Google Scholar 

  5. Dat LD, Tu NTM, Duc NV, Luyen BTT, Huyen CTT, Jang HJ, Thu DT, Huong TT, Tram LH, Thong NV, Hung ND, Kim YH, Thao NP. Anti-inflammatory secondary metabolites from the stems of Millettia dielsiana Harms ex Diels. Carbohydr Res. 2019;484:107778. https://doi.org/10.1016/j.carres.2019.107778.

    Article  CAS  PubMed  Google Scholar 

  6. Rayanil K, Bunchornmaspan P, Tuntiwachwuttikul P. A new phenolic compound with anticancer activity from the wood of Millettia leucantha. Arch Pharm Res. 2011;34(6):881–6. https://doi.org/10.1007/s12272-011-0603-4.

    Article  CAS  PubMed  Google Scholar 

  7. Yenesew A, Derese S, Midiwo JO, Oketch-Rabah HA, Lisgarten J, Palmer R, Heydenreich M, Peter MG, Akala H, Wangui J, Liyala P, Waters NC. Anti-plasmodial activities and X-ray crystal structures of rotenoids from Millettia usaramensis subspecies usaramensis. Phytochem Lett. 2003;64(3):773–9. https://doi.org/10.1016/S0031-9422(03)00373-X.

    Article  CAS  Google Scholar 

  8. Palazzino G, Rasoanaivo P, Federici E, Nicoletti M, Galeffi C. Prenylated isoflavonoids from Millettia pervilleana. Phytochem Lett. 2003;63(4):471–4. https://doi.org/10.1016/S0031-9422(02)00489-2.

    Article  CAS  Google Scholar 

  9. Ngandeu F, Bezabih M, Ngamga D, Tchinda AT, Ngadjui BT, Abegaz BM, Dufat H, Tillequin F. Rotenoid derivatives and other constituents of the twigs of Millettia duchesnei. Phytochem Lett. 2008;69(1):258–63. https://doi.org/10.1016/j.phytochem.2007.05.038.

    Article  CAS  Google Scholar 

  10. Derese S, Barasa L, Akala HM, Yusuf AO, Kamau E, Heydenreich M, Yenesew A. 4’-Prenyloxyderrone from the stem bark of Millettia oblata ssp teitensis and the antiplasmodial activities of isoflavones from some Millettia species. Phytochem Lett. 2014;8:69–72. https://doi.org/10.1016/j.phytol.2014.02.001.

    Article  CAS  Google Scholar 

  11. Bora MM, Deka R, Ahmed N, Kakati DK. Karanja (Millettia pinnata (L.) Panigrahi) seed oil as a renewable raw material for the synthesis of alkyd resin. Ind Crops Prod. 2014;61:106–4. https://doi.org/10.1016/j.indcrop.2014.06.048.

    Article  CAS  Google Scholar 

  12. Jena R, Rath D, Rout SS, Kar DM. A review on genus millettia: traditional uses, phytochemicals and pharmacological activities. Saudi Pharm J. 2020;28(12):1686–93. https://doi.org/10.1016/j.jsps.2020.10.015.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Deyou T, Marco M, Heydenreich M, Pan F, Gruhonjic A, Fitzpatrick PA, Koch A, Derese S, Pelletier J, Rissanen K, Yenesew A, Erdélyi M. Isoflavones and rotenoids from the leaves of Millettia oblata ssp. teitensis. J Nat Prod. 2017;80(7):2060–6. https://doi.org/10.1021/acs.jnatprod.7b00255.

    Article  CAS  PubMed  Google Scholar 

  14. Pancharoen O, Athipornchai A, Panthong A, Taylor WC. Isoflavones and rotenoids from the leaves of Millettia brandisiana. Chem Pharm Bull. 2008;56(6):835–8. https://doi.org/10.1248/cpb.56.835.

    Article  CAS  Google Scholar 

  15. Deyou T, Gumula I, Pang F, Gruhonjic A, Mumo M, Holleran J, Duffy S, Fitzpatrick PA, Heydenreich M, Landberg G, Derese S, Avery V, Rissanen K, Erdelyi M, Yenesew A. Rotenoids, flavonoids, and chalcones from the root bark of Millettia usaramensis. J Nat Prod. 2015;789(12):2932–9. https://doi.org/10.1021/acs.jnatprod.5b00581.

    Article  CAS  Google Scholar 

  16. Chen DL, Liu YY, Ma GX, Zhu NL, Wu HF, Wang DL, Xu XD. Two new rotenoids from the roots of Millettia speciosa. Phytochem Lett. 2015;12:196–9. https://doi.org/10.1016/j.phytol.2015.04.003.

    Article  CAS  Google Scholar 

  17. Panda SK, Padhi L, Leyssen P, Liu M, Neyts J, Luyten W. Antimicrobial, anthelmintic, and antiviral activity of plants traditionally used for treating infectious disease in the similipal biosphere reserve, Odisha, India. Front Pharmacol. 2017;8:1–15. https://doi.org/10.3389/fphar.2017.00658.

    Article  CAS  Google Scholar 

  18. Pulipati S, Babu PS, Lakshmi DN, Navyasri N, Harshini Y, Vyshnavi J, Prasanth M. A phyto pharmacological review on a versatile medicinal plant: Pongamia pinnata (L.) pierre. J Pharmacogn Phytochem. 2018;7(4):459–3. https://www.phytojournal.com/archives/2018/vol7issue4/PartH/7-3-585-478.pdf.

    CAS  Google Scholar 

  19. Huang Z, Zeng YJ, Chen X, Luo SY, Pu L, Li FZ, Zong MH, Lou WY. A novel polysaccharide from the roots of Millettia Speciosa champ: preparation, structural characterization and immunomodulatory activity. Int J Biol Macromol. 2020;145:547–57. https://doi.org/10.1016/j.ijbiomac.2019.12.166.

    Article  CAS  PubMed  Google Scholar 

  20. Rajakumar G, Gomathi T, Thiruvengadam M, Devirajeswari V, Kalpana VN, Chung IM. Evaluation of anticholinesterase, antibacterial and cytotoxic activities of green synthesized silver nanoparticles using from Millettia pinnata flower extract. Microb Pathog. 2017;103:123–8. https://doi.org/10.1016/j.micpath.2016.12.019.

    Article  CAS  PubMed  Google Scholar 

  21. Tu Y, Wu C, Kang Y, Li Q, Zhu C, Li Y. Bioactivity-guided identification of flavonoids with cholinesterase and β-amyloid peptide aggregation inhibitory effects from the seeds of Millettia Pachycarpa. Bioorg Med Chem Lett. 2019;29(10):1194–8. https://doi.org/10.1016/j.bmcl.2019.03.024.

    Article  CAS  PubMed  Google Scholar 

  22. Das S, Ganapaty S. In vitro anthelmintic activity of Millettia auriculata leaves and stems. Asian Pac J Trop Dis. 2014;4(2):S870–3. https://doi.org/10.1016/S2222-1808(14)60749-0.

    Article  Google Scholar 

  23. Tang H, Pei HY, Wang TJ, Chen K, Wu B, Yang QN, Zhang Q, Yang JH, Wang XY, Tang MH, Peng AH, Ye HY, Chen LJ. Flavonois and biphenylneolignans with anti-inflammatory activity from the stems of Millettia griffithii. Bioorg Med Chem Lett. 2016;26(18):4417–22. https://doi.org/10.1016/j.bmcl.2016.08.009.

    Article  CAS  PubMed  Google Scholar 

  24. Sikarwar MS, Patil MB. Antidiabetic activity of Pongamia pinnata leaf extracts in alloxan-induced diabetic rats. Int J Ayurveda Res. 2010;1(4):199–204. https://doi.org/10.4103/0974-7788.76780.

    Article  PubMed  PubMed Central  Google Scholar 

  25. Pailee P, Mahidol C, Ruchirawat S, Prachyawarakorn V. Diverse flavonoids from the roots of Millettia brandisiana. Phytochemistry. 2019;162:157–64. https://doi.org/10.1016/j.phytochem.2019.03.013.

    Article  CAS  PubMed  Google Scholar 

  26. Bruker. APEX3, SAINT-Plus and SADABS. Madison: Bruker AXS Inc.; 2016.

    Google Scholar 

  27. Sheldrick, G.M. (2015) SHELXT-Integrated Space-Group and Crystal-Structure Determination. Acta Crystallographica Section A, A71, 3-8. https://doi.org/10.1107/S2053273314026370.

  28. Sheldrick, G.M. (2015) Crystal Structure Refinement with SHELXL. Acta Crystallographica C, C71, 3-8. https://doi.org/10.1107/S2053229614024218.

  29. Farrugia LJ. WinGX and ORTEP for Windows: an update. J Appl Cryst. 2012;45:849–54. https://doi.org/10.1107/S0021889812029111.

    Article  CAS  Google Scholar 

  30. Perera MMN, Dighe SN, Katavic PL, Collet TA. Antibacterial potential of extracts and phytoconstituents isolated from syncarpia hillii leaves In Vitro. Plants. 2022;11(3):283. https://doi.org/10.3390/plants11030283.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Pompimon W, Sombutsiri, P, Baison W, Udomputtimekakul P, Chuajedton A, Suksen K, Limthongkul J, Naparswad C. Cancer cytotoxic and anti-HIV potential of triphala herb mixture on from Chae Son, Lampang, Thailand. J Pharm Res Int. 2019;1–9. https://journaljpri.com/index.php/JPRI/article/view/30285.

  32. Tan GT, Pezzuto JM, Kinghorn AD, Hughes SH. Evaluation of natural products as inhibitors of human immunodeficiency virus type 1 (HIV-1) reverse transcriptase. J Nat Prod. 1991;54(1):143–54. https://doi.org/10.1021/np50073a012. PMID: 1710653.

    Article  CAS  PubMed  Google Scholar 

  33. Silprasit K, Thammaporn R, Tecchasakul S, Hannongbua S, Choowongkomon K. Simple and rapid determination of the enzyme kinetics of HIV-1 reverse transcriptase and anti-HIV-1 agents by a fluorescence based method. J Virol Methods. 2011;171(2):381–7. https://doi.org/10.1016/j.jviromet.2010.12.001.

    Article  CAS  PubMed  Google Scholar 

  34. Kiser R, Makovsky S, Terpening SJ, Laing N, Clanton DJ. Assessment of a cytoprotection assay for the discovery and evaluation of anti-human immunodeficiency virus compounds utilizing a genetically-impaired virus. J Virol Methods. 1996;58(1–2):99–109. https://doi.org/10.1016/0166-0934(95)01998-7.

    Article  CAS  PubMed  Google Scholar 

  35. Nara PL, Hatch WC, Dunlop NM, Robey WG, Arthur LO, Gonda MA, Fischinger PJ. Simple, Rapid, Quantitative, Syncytium-forming microassay for the detection of human immunodeficiency virus neutralizing antibody. AIDS Res Hum Retrovir. 1987;3:283–302.

    Article  CAS  PubMed  Google Scholar 

  36. Hongthong S, Kuhakarn C, Jaipetch T, Prabpai S, Kongsaeree P, Piyachaturawat P, et al. Polyoxygenated cyclohexene derivatives isolated from Dasymaschalon sootepense and their biological activities. Fitoterapia. 2015;106:158–66.

    Article  CAS  PubMed  Google Scholar 

  37. Skehan P, Storeng R, Scudiero D, Monks A, McMahon J, Vistica D, Warren JT, Bokesch H, Kenney S, Boyd MR. New colorimetric cytotoxicity assay for anticancer-drug screening. J Natl Cancer Inst. 1990;82(13):1107–12. https://doi.org/10.1093/jnci/82.13.1107. PMID: 2359136.

    Article  CAS  PubMed  Google Scholar 

  38. Ahmed M, Shireen KF, Rashid MA, Ul-Ameen M. A further rotenoid from Derris elliptica. Planta Med. 1989;55:207–8.

    Article  CAS  Google Scholar 

  39. Lin YL, Chen YL, Kuo YH. A novel 12-deoxo-12α-acetoxyelliptone, from the roots of Derris oblonga. J Nat Prod. 1993;56(7):1187–9. https://doi.org/10.1021/np50097a029.

    Article  CAS  Google Scholar 

  40. Nakatani N, Matsui M. Synthetic studies on rotenoids. Part I. A novel synthesis of (±)-munduserone. Agr Biol Chem. 1968;32(6):769–2. https://doi.org/10.1080/00021369.1968.10859125.

  41. Trnkova L, Huska D, Adam V, Kizek R, Eckschlager T, Stiborova M, Hubalek J. Electrochemical biosensor for investigation of anticancer drugs interactions (doxorubicin and ellipticine) with DNA. In Proceedings of the SENSORS, Christchurch, New Zealand, 25–28, 2009, 1200–1203.

  42. Stiborová M, Sejbal J, Borek-Dohalská L, Aimová D, Poljaková J, Forsterová K, Rupertová M, Wiesner J, Hudecek J, Wiessler M, Frei E. The anticancer drug ellipticine forms covalent DNA adducts, mediated by human cytochromes P450, through metabolism to 13-hydroxyellipticine and ellipticine N2-oxide. Cancer Res. 2004;64(22):8374–80. https://doi.org/10.1158/0008-5472.CAN-04-2202. PMID: 15548707.

    Article  PubMed  Google Scholar 

  43. Hofmayer MS, Sunagatullina A, Brösamlen D, Mauker P, Knochel P. Stereoselective Cobalt-Catalyzed Cross-Coupling Reactions of Arylzinc Chlorides with α-Bromolactones and Related Derivatives. Organic Letters. 2020;22(4):1286–9. https://doi.org/10.1021/acs.orglett.9b04564.

    Article  CAS  Google Scholar 

  44. Ahmad-Junan SA, Amos PC, Whiting DA. Novel and efficient synthesis of rotenoids via intramolecular radical arylation. Journal of the Chemical Society, Perkin Transactions 1. 1992(5):539–45.

  45. Onyegeme-Okerenta B, Okafor U. Antimicrobial Properties of Ethanol Leaf Extract of Millettia aboensis on Some Selected Clinical Isolates. Univers J Plant Sci. 2014;2:97–101.

    Article  Google Scholar 

  46. Egyir B, Baffour-Awuah Owusu K, Ohashi M, Mary Anti Chama MA. Phytochemical composition and In vitro antibacterial activities of Millettia chrysophylla and Milletia zechiana. J Sci Technol. 2022;40(1):66–85.

    Google Scholar 

  47. Panthong K, Hongthong S, Kuhakarn C, Piyachaturawat P, Suksen K, Panthong A, et al. Pyranonaphthoquinone and anthraquinone derivatives from Ventilago harmandiana and their potent anti-inflammatory activity. Phytochemistry. 2020;169:112182.

    Article  CAS  PubMed  Google Scholar 

  48. Sangsuwon C, Jiratchariyakul W. Antiproliferative Effect of Lung Cancer Cell Lines and Antioxidant of Macluraxanthone from Garcinia Speciosa Wall. Procedia Soc Behav Sci. 2015;197:1422–7.

    Article  Google Scholar 

  49. Kim SH, Song GY, Sok DE, Ahn BZ. Anti-cell adhesive effect of phenylacetylshikonin analogues related to their cytotoxicity in A549 cells. Arch Pharm Res. 1997;20(2):155–7. https://doi.org/10.1016/j.phytochem.2019.1121827.doi:10.1007/BF02974003. PMID: 18975194.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

This research project is supported by the Center of Excellence for Innovation in Chemistry (PERCH-CIC), Ministry of Higher Education, Science, Research and Innovation and is gratefully acknowledged. The authors thank Dr. Banchob Sripa from Liver Fluke and Cholangiocarcinoma Research Center, Department of Pathology, Faculty of Medicine, Khon Kean University for KKU-M213 cells. Furthermore, we are grateful to the Department of Microbiology for AID examination and the Department of Chemistry, Mahidol University for NMR and MS measurements.

Funding

This research was not funded by any funding source.

Author information

Authors and Affiliations

Authors

Contributions

SS, AK, WP, and PU conceived and designed all the experiments, performed isolation and identification of compounds, drafted and reviewed the manuscript. SN, SJ, and KC carried out the X-ray analysis section. NN, BW, and NC prepared and identified plant material. KS, AC, JL, and CN performed the bioactivity evaluation section. SP, and PM carried out the spectroscopic analysis section. All authors have read and approved the manuscript.

Corresponding author

Correspondence to Phansuang Udomputtimekakul.

Ethics declarations

Ethics and consent to participate

In this study, the plant extracts and pure compounds were tested for their antibacterial, anti-HIV viral activity, and inhibitory effect on cancer cells. The research team conducted the in vitro testing according to the methods approved by the Ethics Committee of the Faculty of Science, Mahidol University.

Consent for publication

Not applicable.

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Sananboonudom, S., Kaewnoi, A., Pompimon, W. et al. Study on the absolute configuration and biological activity of rotenoids from the leaves and twigs of Millettia pyrrhocarpa Mattapha, Forest & Hawkins, sp. Nov. BMC Complement Med Ther 23, 147 (2023). https://doi.org/10.1186/s12906-023-03963-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12906-023-03963-4

Keywords