Next Article in Journal
Chemical Constituents and Bioactivities of the Plant-Derived Fungus Aspergillus fumigatus
Previous Article in Journal
Synthesis and Evaluation of Antimicrobial Activity of the Rearranged Abietane Prattinin A and Its Synthetic Derivatives
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Unusual Vilasinin-Class Limonoids from Trichilia rubescens

1
School of Pharmaceutical Sciences, College of Medical, Pharmaceutical and Health Sciences, Kanazawa University, Kanazawa 920-1192, Japan
2
School of Pharmacy, Tokyo University of Pharmacy and Life Sciences, Hachioji 192-0392, Japan
3
Natural Products Branch, Developmental Therapeutics Program, Center of Cancer Research, Division of Cancer Treatment and Diagnosis, National Cancer Institute, Frederick, MD 21702-1201, USA
4
Molecular Targets Program, Center for Cancer Research, Division of Cancer Treatment and Diagnosis, National Cancer Institute, Frederick, MD 21702-1201, USA
5
Natural Products Research Laboratories, UNC Eshelman School of Pharmacy, University of North Carolina at Chapel Hill, Chapel Hill, NC 27599-7568, USA
6
Chinese Medicine Research and Development Center, China Medical University and Hospital, 2 Yuh-Der Road, Taichung 40447, Taiwan
*
Author to whom correspondence should be addressed.
Deceased 24 October 2021.
Molecules 2024, 29(3), 651; https://doi.org/10.3390/molecules29030651
Submission received: 5 January 2024 / Revised: 17 January 2024 / Accepted: 23 January 2024 / Published: 30 January 2024
(This article belongs to the Section Natural Products Chemistry)

Abstract

:
Eight vilasinin-class limonoids, including the unusually chlorinated rubescins K–M (13), the 2,3-epoxylated rubescin N (4), and rubescins O–R (58), were newly isolated from Trichilia rubescens. The structures of the isolated compounds were determined through spectroscopic and spectrometric analyses, as well as ECD calculations. The natural occurrence of chlorinated limonoids 13 was confirmed by chemical methods and HPLC analysis of a roughly fractionated portion of the plant extract. Eight selected limonoids, including previously known and new compounds, were evaluated for antiproliferative activity against five human tumor cell lines. All tested limonoids, except 8, exhibited significant potency, with IC50 values of <10 μM; in particular, limonoid 14 strongly inhibited tumor cell growth, with IC50 values of 0.54–2.06 μM against all tumor cell lines, including multi-drug-resistant cells.

Graphical Abstract

1. Introduction

Tropical rainforests are renowned for their extraordinary biodiversity, which is the result of intricate ecosystems, a diverse array of species, and extensive genetic diversity within those species. As part of ongoing research focused on the phytochemical exploration of tropical rainforests, we investigated Trichilia rubescens (Meliaceae), which displayed significant antiproliferative activity in an NCI-60 panel screening (Figure S1 in the Supplementary Materials).
The family Meliaceae has approximately 50 genera and about 1400 species; however, only limited genera have been investigated phytochemically [1]. The genus Trichilia, commonly found in the tropical and subtropical regions of South America and Africa [2], is a well-studied genus known to produce various limonoids [3]. Limonoids are a structurally unique class of tetra-nor-triterpenes containing a furan ring and exhibit many biological activities, such as antitumor, anti-HIV, insecticidal, and anti-inflammatory activities [3,4]. Interestingly, limonoids have been isolated from specific restricted families, primarily Meliaceae and Rutaceae, with approximately 2700 meliaceous limonoids being isolated and identified to date [1]. Among the genera in the former family, the genus Trichilia yields ring-intact limonoids in a high 25% ratio, while other genera produce mostly rearranged limonoids [1]. Vilasinin is a ring-intact limonoid characterized by a tetrahydrofuran ring adjacent to the A and B rings (Figure 1). Currently, only 18 vilasinin-class limonoids have been reported from the genus Trichilia, with 15 isolated from T. rubescens [5,6,7,8,9,10,11,12].
In this study, extensive phytochemical research on T. rubescens led to the identification of eight new vilasinin-class limonoids, including the unusually chlorinated rubescins K–M (13), the 2,3-epoxylated rubescin N (4), and rubescins O–R (58) (Figure 2), along with six known vilasinin-class limonoids, rubescin E (9) [6], rubescin F (10) [7], rubescin H (11) [7], rubescin I (12) [8], rubescin J (13) [8], and TS3 (14) [9], three sesquiterpenes, guaianediol [13], alismol [14], and eudesm-4(15)-ene-1β,6α-diol [15], as well as two sterols, β-sitosterol and stigmasterol [16] (Figure 3). The natural occurrence of the chlorinated limonoids 13 was further confirmed using chemical methods and HPLC analysis. This paper deals with the isolation, structure elucidation, and evaluation of antiproliferative activity against human tumor cell lines of the isolated limonoids.

2. Results and Discussion

A 50% MeOH/CH2Cl2 extract (N047159) of the leaves of T. rubescens was partitioned with EtOAc and water. The EtOAc-soluble fraction was subjected to a series of chromatographic techniques using silica gel and octadecylsilica (ODS) gel via medium-pressure liquid chromatography (MPLC), column chromatography (CC), preparative TLC (pTLC), and HPLC to obtain pure new limonoids 18, along with the known compounds 914. The structures of all known compounds were identified by a comparison of their 1D NMRs with previously reported values [6,7,8,9].
Compound 1 was isolated as a colorless amorphous solid with a specific rotation of α D 21   −22 (c 0.1, CHCl3). The HRAPCIMS spectrum of 1 showed protonated molecular peaks at m/z 595.2118 and 597.2100 [M+H]+ (calcd 595.2099 and 597.2069) in a 3:1 ratio, which suggested the presence of chlorine and a molecular formula of C33H35ClO8. The IR absorptions at 3257 and 1720 cm−1 implied the presence of hydroxy and carbonyl groups, respectively. The 1H NMR spectrum showed signals assignable to protons for a mono-substituted phenyl (δH 7.92, 7.62, 7.47), three olefinic, three methylene, eight methine, and four methyl groups (Table 1). The 13C NMR spectrum of 1 (Table 2) presented 33 peaks, including the signals for a ketone carbonyl (δC 199.4), an ester carbonyl (δC 165.9), four olefinic, and seven tertiary carbons. These signals were similar to those of TS2 (Figure S66 in the Supplementary Materials), a vilasinin-class limonoid with two β-epoxy groups at C-9, -11 and C-14, -15, a methacrylate at C-7, and an α,β-unsaturated ketone in ring-A, which was isolated previously from the leaves of T. rubescens [9]. However, the chemical shifts of C-2, C-3, and the ester moiety at C-7 in 1 differed significantly from those in TS2; therefore, compound 1 likely lacks a C-2, C-3 double bond and has a different ester group at C-7. The planar structure of 1 was further confirmed by 2D NMR experiments (Figure 4). 1H–1H COSY correlations were observed between OH-3/H-3, H-2/H-3, H-5/H-6/H-7, H-11/H-12, H-16/H-17, and H-22/H-23. The HMBC correlations from H3-19 to C-1 and H-2 to C-1 suggested a carbonyl at C-1, while the correlations from H3-29 to C-3 indicated a hydroxy group at C-3. A benzoyl ester was assigned at C-7, based on the HMBC correlations of H-7 and H-3′ with C-1′. The positions of the two epoxy groups at C-9, C-11 and C-14, C-15 were confirmed, based on multiple HMBC cross-peaks from H3-18, H2-16, and H3-30 to oxygenated carbons at C-14, C-15, and C-9, respectively. A pendant furan ring at C-17 was suggested by HMBC correlations from H-17 to C-20 and C-22, as well as from H-23 to C-20 and C-21. These data indicate that compound 1 has the same basic skeleton as TS2, but has a chlorine at C-2, a hydroxy at C-3, and a benzoate rather than methacrylate at C-7. The key NOESY correlations (Figure 5) from H-2 to H3-19/H3-29, H-6 to H-7/H3-19/H3-29, H-7 to H-15/H3-30, H-15 to H3-30, H-12α to H3-18/H-11, and H-12β to H-17 suggested α-orientations for the chlorine at C-2, the methyl at C-18, the ester carbonyl at C-7, and the furan ring at C-17; the latter three assignments were consistent with those of other related vilasinin-class limonoids, such as TS2 [9]. The absolute configuration of 1 was deduced by comparing the experimental and calculated ECD spectra (Figure 6). Thus, the structure of 1 (rubescin K) was elucidated as the first chlorinated vilasinin-class limonoid, with (2S, 3R, 4R, 5S, 6R, 7S, 8S, 9S, 10S, 11S, 13S, 14R, 15R, 17S) absolute configurations.
Compounds 2 (rubescin L) and 3 (rubescin M) were obtained as optically active colorless amorphous solids. Their HRAPCIMS spectra also indicated the presence of chlorine, and the protonated molecule peaks at m/z 573.2253/575.2231 [M+H]+ (3:1, calcd 573.2255/575.2226) for compound 2 and m/z 621.2248/623.2237 [M+H]+ (3:1, calcd 621.2255/623.2226) for compound 3 indicated the molecular formulae of C31H37ClO8 and C35H37ClO8, respectively. The similarities among the 1D (Table 1 and Table 2) and 2D (Figure 3 and Figure 4) NMR spectra of compounds 13 strongly suggested that all three compounds are 2-chloro-3-hydroxy vilasinin limonoids. The only differences were found in the signals assignable to the C-7 ester moiety. The 13C and 1H NMR spectra of compound 2 suggested a tigloyl group, with an olefinic methine and two methyl carbons, as well as an olefinic and two sets of methyl protons assigned to the ester moiety. Additionally, an HMBC correlation between H-3′ and C-1′ supported the presence of a tigloyl ester, and a good comparison was found with the NMR data of the related tigloyl limonoid, compound 9 [6]. The ester moiety of 3 was elucidated as a cinnamate, based on NMR observations of five phenyl methine and two olefinic protons, as well as HMBC correlations from olefinic protons H-2′ and H-3′ to phenyl carbons C-4′ and C-5′. Although limited isolations of related cinnamoyl limonoids from the Meliaceae have been reported, the comparisons of the NMR data for the ester moiety of compound 3 with those of other cinnamoyl limonoids, such as toosendansins E and F [17], 7-cinnamoyltoosendanin [18], and ohchininolide [19] and its derivatives [20], also supported the presence of a cinnamate at C-7 in 3. Unfortunately, the instability of compounds 2 and 3 did not allow further investigation, such as with UV, IR, and ECD; however, all other data strongly supported that, like 1, both compounds are interesting chlorinated limonoids.
Compound 4 was isolated as a colorless amorphous solid with an optical rotation of α D 21   −13 (c 0.1, CHCl3). With a protonated molecular peak at m/z 455.2070 [M+H]+ (calcd. 455.2064) in its HRFABMS spectrum, the molecular formula of 4 is C26H30O7. The 1H and 13C NMR data (Table 3 and Table 4) of 4 were like those of 10 (Figure 3) [7], a vilasinin-class limonoid with an oxetane ring formed by an ether linkage between C-7 and C-14. However, the signals that were due to the olefinic protons (H-2/H-3) in 10 were replaced by signals for protons (δH 3.15 and δH 7.57) on oxygenated carbons in 4, with a 1H–1H COSY correlation between the two protons. The 13C NMR signals for C-2 and C-3 in 4 (δc 51.9 and δc 60.4) were shielded from those in 10 (δc 131.3 and δc 152.7) [7]. The above data together with HMBC correlations between H-2 and C-4/C-10, H-3 and C-4/C-5, and H3-29 and C-3 indicated the presence of an epoxy ring at C-2/C-3. The relative configurations of 4 were established from a NOESY experiment (Figure 5). The β-orientation of the 2,3-epoxy moiety was suggested by the key correlations of H-3/H2-28α and H2-28α/H-5. The NOESY correlations from H-7 to H-15/H3-30 and from H-15 to H3-30 also indicated a β-orientation for OH-15. A strong NOE correlation was observed between β-oriented H-7 and α-oriented H-15; both protons are pseudo-equatorial, with a proton-proton distance of 2.4 Å. The same relative configurations were also supported by a comparison with all NMR data for compound 10 [7]. The absolute configuration was supplied from the calculated and experimental ECD spectra (Figure 6). Thus, compound 4 (rubescin N) is (2R, 3R, 4R, 5S, 6R, 7S, 8S, 9S, 10S, 11S, 13S, 14S, 15R, 17S)-2,3-epoxyrubescin N, the first reported 2,3-epoxylated vilasinin-class limonoid. Furthermore, it is the second reported vilasinin-class limonoid with an oxetane ring [7], although several other classes of oxetane limonoids have been isolated [21,22,23,24]. As a biogenetic oxetane formation has been proposed previously [21,23], compound 4 might be produced from TS1 (Figure S66), which was also isolated from the same plant [9] through 2,3-epoxidation and the intramolecular nucleophilic attack of OH-7 on C-14 to open the epoxy ring.
Compound 5 (rubescin O) has a molecular formula of C33H34O7, as indicated by HRFABMS. Its 1H and 13C NMR signal patterns were nearly identical to those of compound 1, suggesting a vilasinin skeleton for 5 with two epoxy moieties at C-9 (δC 64.6)/C-11 (δC 60.4/δH 3.98) and C-14 (δC 68.3)/C-15 (δC 55.2/δH 3.50), a furan ring at C-17 (δC 38.7/δH 2.47), and a benzoate at C-7 (δC 75.5/δH 5.68). However, significantly different chemical shifts were observed at C-2 (δC 131.3/δH 5.99) and C-3 (δC 151.7/δH 7.10) (Table 3 and Table 4), which implied the presence of an olefin creating an α,β-unsaturated ketone with C-1 (δC 200.3). This assignment was confirmed via comparisons with the NMR data of the related limonoids with an α,β-unsaturated ketone, such as compound 9 (Figure 3) [6]. All 2D NMR, including COSY, HMBC, and NOESY, supported the planar structure and the relative configurations of 5 (Figure 4 and Figure 5). The absolute configurations of 5 (4R, 5S, 6R, 7S, 8S, 9S, 10S, 11S, 13S, 14R, 15R, 17S) were deduced via a comparison of the experimental and calculated ECD spectra (Figure 6).
The molecular formula of compound 6 (rubescin P) was elucidated as C28H34O6, based on HRFABMS. The 1D NMR data for 6 were related to those of rubescins C [5] and J [8], which are vilasinin-class limonoids with an acetate at C-11. The major difference was the chemical shift at C-7 (δC 75.4) and the appearance of a proton at δH 4.09, suggesting the presence of a hydroxy group at C-7 (Table 3 and Table 4). The HMBC correlations of H-11/C-1′ and H3-2′/C-1′ (Figure 4) and the NOESY correlations between H-11 and H3-19/H-12β provided further support for an α-oriented acetoxy group at C-11 (Figure 5). The NOESY correlations of H-7/H-6/H3-30 indicated that the C-7 hydroxy group was β-oriented. The absolute configurations of (4R, 5S, 6R, 7S, 8R, 9S, 10R, 11S, 13S, 17R) were deduced from the calculated and experimental ECD spectrum (Figure 6).
Compound 7 (rubescin Q) has a molecular formula of C26H28O7, as suggested by HRFABMS. The 1D and 2D NMR spectroscopic data (Table 3 and Table 4) of 7 indicated a comparable structure to that of 11 [7], in which a cyclopropane ring is formed between C-7, C-8, and C-14. However, the pendant furan ring found at C-17 in 11 was not present in 7. Instead, a 5-hydroxy-2-oxo-dihydrofuran ring was present in 7, based on the HMBC correlations (Figure 4) from H-17 to C-20 (δC 138.8/138.9) and from H-23 to C-20 (δC 138.8/138.9) and H-22 to C-21 (δC 171.3/171.0), together with a 1H–1H COSY correlation between H-22 and H-23. Also, the related chemical shifts were identical to those of trichirubine A, which has the same substituent at C-17 [10]. The 1:1 pairwise signals observed in the 1H and 13C NMR spectra indicated that 7 was a diastereomixture caused by the epimerization of a hemiacetal moiety at C-23. The same observation was also made with rubescin G [7] and trichirubine A [10]. Based on the relative configurations revealed by NOESY correlations (Figure 4), the NOESY cross-peaks of H-7/H3-30/H-15 confirmed that a hydroxy group at C-15 and a proton at C-7 were β-oriented. The rigid conformation around the cyclopropane ring necessarily allowed the construction of the R configuration at C-14. The calculated and experimental ECD (Figure 6) elucidated the absolute configuration of 7 as (4R, 7S, 8S, 9S, 10S, 11S, 13S, 14R, 15R, 17R).
HRFABMS suggested a molecular formula of C26H30O6 for compound 8 (rubescin R). A comparison of the NMR data (Table 3 and Table 4) of 8 with those of compound 11 [7] (Figure 3) showed overall similarity, except at C-5, C-6, C-7, C-28, and C-29. The most significant differences were the shielded C-5 (δc 55.4) and the deshielded C-6 (δc 207.1) and C-7 (δc 44.4) in 8 compared with 11, together with the appearance of a second hydroxy proton (δH 1.64, 28-OH) and an aliphatic methine proton (δH 3.67, H-5). These data suggested the absence of a C-5, C-6 double bond, the presence of a carbonyl at C-6, and the cleavage of the ether linkage between C-6 and C-28. The HMBC correlations of H-5 and H-7 with C-6 also supported the presence of a carbonyl at C-6 (δC 207.1), while the shielded carbon signal for C-28 (δC 70.3), the COSY connectivity of H2-28/OH-28, and the HMBC cross-peaks of H2-28 to C-4/C-5 indicated the presence of a hydroxymethyl moiety at C-4. While a NOESY analysis (Figure 5) confirmed the relative configuration of 8, an ECD analysis (Figure 6) and the negative specific rotation value determined that 8 and 11 have the same absolute configuration (4R, 5S, 7S, 8S, 9S, 10S, 11S, 13S, 14R, 15R, 17R). All the above data indicate that 8 is a seco-vilasinin-class limonoid and is a possible biosynthetic intermediate of 11.
A possible biosynthetic pathway to the chlorinated limonoids 13 could involve the epoxidation of related limonoids, such as 5, 9, and 15, which contain an α,β-unsaturated ketone in ring-A to give the tri-epoxy 16, followed by chlorination at C-2 (Figure 7). Limonoids 5 and 9 were isolated in this study. While halogenation is thought to primarily be a final biosynthetic step, limited evidence has indicated that halogenation can also be followed by subsequent steps that lead to other biosynthetic intermediates [25]. Accordingly, there is a slight possibility that an epoxy group might be produced by the attack of a hydroxy group and the removal of a neighboring chlorine atom.
Halogenated natural products are produced mainly by marine organisms living in halogen-rich environments, or by microorganisms such as algae, cyanobacteria, and fungus [26,27,28,29,30,31]. Although rarely isolated from terrestrial plants, several plant-derived natural products containing halogen, usually chlorine, have been reported [32,33,34,35]. However, some of these compounds could be artifacts that are produced when halogenated solvents are used. Since this study is the first to report chlorinated limonoids and halogen-containing natural products from the family Meliaceae, further investigation was needed to determine whether they occur naturally in the plant extract.
Chlorination of the related β-2,3-epoxy limonoid 16 (Figure 7) could happen during the isolation process, for example, with the use of CHCl3 or CH2Cl2 under acidic conditions, such as on silica gel, or unexpectedly, by contamination with HCl. To investigate the possibility that the isolated chlorinated limonoids might be artifacts, the model substrate 17 was prepared as a biosynthetic precursor mimic of 1. The known limonoid 11, which was isolated in sufficient quantities in this study, was epoxidized via a general condition using H2O2 to give 17 at a 68% yield (Figure 8) [36]. The β-orientation of the 2,3-epoxide was confirmed by a NOESY correlation between H-3 and H-28α, which was also seen with 4. The treatment of 17 with excess silica gel in CHCl3 for 12 h produced no reaction; only the starting material was recovered. This observation suggested a low probability of chlorination occurring during silica gel column chromatography with a chlorinated solvent. The treatment of 17 with dried HCl on silica gel in CHCl3 gave a complex inseparable mixture rather than a chlorinated product [37]. The 1H-NMR spectrum of the mixture showed no characteristic peaks at H-2 and H-3 corresponding to the 2-chloro-3-hydroxy segment in 1. This result indicated that the chlorination of a related epoxide was unlikely, even after unexpected contamination with HCl.
To further confirm that the isolated chlorinated limonoids are natural products, LC/MS analysis was carried out on the initial fraction 7b containing 1, which was roughly twice separated from the original extract using silica gel CC. The peak with the exact mass for an authentic sample of 1 was observed within a range of 4.3–4.5 min, and the same peak was detected in fraction 7b at the same retention time (Figures S64 and S65 in the Supplementary Materials).
Based on the results of the above chemical methods and LC/MS analysis, it is highly possible that chlorinated limonoids 13 occur naturally in T. rubescens.
Eight selected vilasinin-class limonoids (1, 5, 6, 8, 9, 11, 13, and 14) were evaluated for their antiproliferative activities against human tumor lines (HTCLs), A549 (lung adenocarcinoma), MDA-MB-231 (triple-negative breast cancer), MCF-7 (HER2-negative), KB (HeLa derivative), and KB-VIN [P-gp overexpressing multidrug-resistant (MDR) KB subline] (Table 5). All the tested limonoids, except 8, showed significant activity, with IC50 values of 0.54–8.46 μM, even against the KB-VIN MDR cell line, which suggested that these limonoids are not P-gp substrates. In particular, compound 14 exhibited the highest potency against all HTCLs, including the MDR tumor cell line. Compounds 9 and 14 have also been reported to effectively decrease the viability of hepatoma cells at TC50 concentrations ranging from 5 to 10 μM. [38]. Comparing the results of the current study, these two compounds show potential value as anticancer drugs. The chlorine-containing limonoid 1 showed slightly selective inhibition against the MCF-7 cell line (IC50 1.34 μM) compared with the other tested HTCLs (IC50 2.53–5.72 μM). Compounds 1 and 5 showed similar activity, indicating that the 2,3-double bond is not very important for this activity. The comparison of 5, 9, and 14 suggested that the function group at C-7 has an insignificant effect on this activity, although compound 14, with a 6,7-double bond, displayed slightly more potent activity than other compounds. Since compound 8 did not exhibit significant antiproliferative activity (IC50 > 40 μM), the tetrahydrofuran ring, which is formed by an ether linkage between C-6 and C-28, might be important for the antiproliferative activity of this compound type [39].

3. Materials and Methods

3.1. General Experimental Procedures

Optical rotations were determined using a JASCO P-2200 digital polarimeter (JASCO, Tokyo, Japan). The UV spectra were measured with a Thermo Scientific Genesys 50 spectrometer (Thermo Scientific, Waltham, MA, USA). ECD spectra were measured with a JASCO J-820 spectrometer. Infrared spectra (IR) were recorded with a Shimadzu IRSpirit FT-IR spectrometer (Shimadzu, Kyoto, Japan) with a QATR-S single reflection ATR accessory, using neat samples. The NMR spectra were obtained via the JEOL JMN-ECA600 and JMN-ECS400 NMR spectrometers (JEOL, Akishima, Japan), with tetramethylsilane as an internal standard; chemical shifts are stated as δ values. HRMS data were recorded on a JMS-700 MStation (FAB) mass spectrometer. MPLC analysis was performed with C18 cartridges (Sfär C18 D, Biotage, Charlotte, NC, USA). Preparative HPLC analysis was conducted on a GL Science recycling system using an InertSustain C18 column (5 μm; 20 × 250 mm) (GL Sciences, Tokyo, Japan). LC/MS analysis was recorded with a Shimadzu LCMS-9030 connected to an Imtakt Cadenza CX-C18 (2 × 50 mm; 3 μm) (Imtakt, Kyoto, Japan).

3.2. Plant Material

A 50% CH2Cl2/MeOH extract of T. rubescens leaves (N047159) was provided by the NCI Natural Products Branch (Developmental Therapeutics Branch, Frederick, MD, USA) as reported previously [40]. The plants were collected in the Central Africa Republic by J. M. Fay in August 1988 and were identified by the taxonomist R. Gereau.

3.3. Extraction and Isolation

The extract N047159 (10.0 g) was partitioned between H2O and EtOAc to obtain H2O-soluble and EtOAc-soluble portions. The EtOAc-soluble portion (5.2 g) was subjected to silica gel CC using gradient mixed solvents [n-hexane/EtOAc (9:1, 8:2, 2:1, 0:1), CH3OH] as eluents to afford eight fractions (frs. 1–8). Fr. 7 was separated by silica gel CC (CHCl3 /MeOH, 1:0–9:1) to give six fractions (frs. 7a–f). Fr. 7a was loaded on ODS-MPLC (MeOH/H2O, 7:3) to give eight subfractions (frs. 7a1–8). Frs. 7a3 and 7a4 were crystallized from MeOH, to give compounds 11 (28.5 mg) and 14 (47.8 mg), respectively. Further separation of Fr. 7a5 by silica gel CC (CHCl3/MeOH, 100:0–95:5), following ODS-HPLC (CH3CN/H2O, 65:35), afforded compounds 13 (24.4 mg), 1 (9.9 mg), and 6 (2.0 mg). Fr. 7a6 was subjected to silica gel CC (CHCl3/MeOH, 100:0–95:5) to give four fractions (frs. 7a6a–d). Frs. 7a6a and 7a6b were purified by ODS-HPLC (MeCN/H2O, 65:35) to yield compounds 5 (3.2 mg) and 9 (1.2 mg). The sequential CC of fr. 7a7 with silica gel CC (n-hexane/EtOAc, 4:1–6:4) and pTLC (n-hexane/EtOAc, 4:1) afforded compound 12 (0.8 mg).
Fr. 7b was subjected to ODS-MPLC (MeOH/H2O, 70:30) to yield five fractions (frs. 7b1–5). Fr. 7b1 was separated by silica gel CC (CHCl3/MeOH, 100:0–95:5) to afford seven fractions (frs. 7b1a–g). Sequential silica gel chromatography (n-hexane/EtOAc, 4:1–1:1 and CHCl3/MeOH, 100:0–99:1) afforded compound 4 (0.7 mg). Fr. 7b2 was separated by silica gel CC (CHCl3/MeOH, 100:0–95:5) and ODS-HPLC (MeCN/H2O, 60:40) to give compounds 1 (3.1 mg) and 2 (2.1 mg). Fr. 7b3 was subjected to silica gel CC (CHCl3/MeOH, 100:0–95:5) to give four fractions (frs. 7b3a–d). Purification of fr. 7b3b by ODS-HPLC (MeCN/H2O, 60:40) furnished compound 1 (18.1 mg). Fr. 7b4 was separated by silica gel CC (CHCl3/MeOH, 100:0–97:3, followed by toluene/EtOAc, 1:0–4:1) and ODS-HPLC (MeCN/H2O, 60:40) to obtain compound 3 (3.4 mg).
Fr. 7c was separated by ODS-MPLC (MeOH/H2O, 60:40) to yield four fractions (frs. 7c1–4). Fr. 7c2 was applied to ODS-MPLC (MeOH/H2O, 40:60) to furnish eight fractions (frs. 7c2a–h). Fr. 7c2b was further separated by ODS-HPLC (MeCN/H2O, 40:60) and silica gel CC (n-hexane/EtOAc, 2:1–1:2, toluene/EtOAc, 3:2–0:1 and then n-hexane/EtOAc, 2:1–1:2) to give compound 10 (0.7 mg).
Fr. 7d was applied to ODS-MPLC (MeOH/H2O, 40:60) to give three fractions (frs. 7d1–3). Fr. 7d2 was separated by ODS–MPLC (MeCN/H2O, 40:60), silica gel CC (CHCl3/MeOH, 100:0–97:3), and ODS-HPLC (MeCN/H2O, 40:60), following silica gel CC (n-hexane/EtOAc, 6:4–4:6), to give compound 8 (3.3 mg).
Fr. 7e was loaded on ODS-MPLC (MeOH/H2O, 70:30 and MeOH/H2O, 50:50) following a silica gel CC (CHCl3/MeOH,100:0–95:5) to give compound 7 (1.2 mg).
Fr. 4 was separated by ODS-MPLC (MeOH/H2O, 70:30 and MeCN/H2O, 70:30) to yield guaianediol (27.5 mg). Fr. 5 was fractionated by silica gel CC (CHCl3/MeOH, 100:0–95:5 and n-hexane/acetone, 9:1–0:1) to afford a mixture of β-sitosterol and stigmasterol (13.2 mg). Fr. 6 was loaded onto a Sephadex LH-20 column (MeOH/H2O, 60:40–90:10), silica gel CC (n-hexane/CH2Cl2/EtOAc, 5:4:1, 5:3:2, 0:0:1 and CHCl3/MeOH, 100:0–98:2) following ODS-HPLC (MeCN/H2O, 40:60) to yield eudesm-4(15)-ene-1β,6α-diol (1.2 mg).
Rubescin K (1): Colorless amorphous solid; α D 21   −22 (c 0.1, CHCl3); UV (MeOH) λmax (logε) 226 (4.04); ECD (MeCN) λmax (Δε) 222 (+1.92); IR (neat) νmax 3857, 2941, 2870, 1720, 1278 cm−1, HRAPCIMS m/z 595.2118 [M+H]+ (calcd for C33H36ClO8, 595.2099); 1H and 13C NMR data (Table 1 and Table 2).
Rubescin L (2): Colorless amorphous solid; α D 21   −25 (c 0.1, CHCl3); HRAPCIMS m/z 573.2253 [M+H]+ (calcd for C31H38ClO8, 573.2255), 1H and 13C NMR data (Table 1 and Table 2).
Rubescin M (3): Colorless amorphous solid; α D 21   −20 (c 0.1, CHCl3); HRAPCIMS m/z 621.2248 [M+H]+ (calcd for C35H38ClO8, 621.2255), 1H and 13C NMR data (Table 1 and Table 2).
Rubescin N (4): Colorless amorphous solid; α D 20   −13 (c 0.1, CHCl3); UV (MeOH) λmax (logε) 202 (3.85); ECD (MeCN) λmax (Δε) 221 (+6.81); IR (neat) νmax 2925, 1713, 1453, 1378 cm−1; HRFABMS m/z 455.2064 [M+H]+ (calcd for C26H31O7, 455.2070); 1H and 13C NMR data (Table 3 and Table 4).
Rubescin O (5): Colorless amorphous solid; α D 23   +16 (c 0.1, CHCl3); UV (MeOH) λmax (logε) 226 (4.14); ECD (MeCN) λmax (Δε) 222 (+8.50); IR (neat) νmax 2939, 2923, 1720, 1675, 1453 cm−1; HRFABMS m/z 543.2378 [M+H]+ (calcd for C33H35O7, 543.2383); 1H and 13C NMR data (Table 3 and Table 4).
Rubescin P (6): Colorless amorphous solid; α D 23   +33 (c 0.1, CHCl3); UV (MeOH) λmax (logε) 204 (4.02); ECD (MeCN) λmax (Δε) 222 (+5.69); IR (neat) νmax 2926, 1732, 1675 cm−1; HRFABMS m/z 467.2440, [M+H]+ (calcd for C28H35O6, 467.2434); 1H and 13C NMR data (Table 3 and Table 4).
Rubescin Q (7): Colorless amorphous solid; α D 20   −81 (c 0.1, CHCl3); UV (MeOH) λmax (logε) 218 (3.98); ECD (MeCN) λmax (Δε) 223 (–7.77); IR (neat) νmax 3468, 2930, 1752, 1723, 1700, 1655 cm−1; HRFABMS m/z 453.1895 [M+H]+ (calcd for C26H29O7, 453.1913); 1H and 13C NMR data (Table 3 and Table 4).
Rubescin R (8): Colorless amorphous solid; α D 20 −29 (c 0.1, CHCl3); UV (MeOH) λmax (logε) 216 (4.14); ECD (MeCN) λmax (Δε) 220 (+11.89); IR (neat) νmax 3468, 2926, 1672 1665 cm−1; HRFABMS m/z 439.2109 [M+H]+ (calcd for C26H31O6, 439.2121); 1H and 13C NMR data (Table 3 and Table 4).
Preparation of compound 17: 5% NaOH (0.2 mL) and 30% H2O2 (0.099 mL) were added to compound 11 (17.0 mg) in MeOH (1.0 mL) and the solution was stirred at room temperature for 11.5 h. The completion of the reaction was confirmed with TLC. The reaction mixture was quenched with the 1M HCl (20 mL). The resulting mixture was then extracted 3 times with EtOAc. The organic layer was dried over Na2SO4, filtered, and evaporated under vacuum [36]. The crude product was purified via silica gel CC (n-hexane: EtOAc; 1:1) to yield compound 17 as a colorless amorphous solid (12.0 mg).
2,3-Epoxyrubescin H (17): Colorless amorphous solid; α D 20 −281 (c 0.1, CHCl3), UV (MeOH) λmax (logε) 206 (4.12); IR (neat) νmax 2925, 1713, 1453, 1378 cm−1, HRFABMS m/z 437.1962 [M+H]+ (calcd for C26H29O6, 437.1964), 1H NMR (600 MHz, CDCl3) 7.38 (1H, t, J = 1.6 Hz, H-23), 7.22 (1H, dd, J = 1.6, 0.7 Hz, H-21), 6.27 (1H, dd, J = 1.6, 0.7 Hz, H-22), 4.42 (1H, d, J = 9.3 Hz, H-28a), 4.27 (1H, d, J = 9.3 Hz, H-28b), 3.9 (1H, brs, H-15), 3.4 (1H, dd, J = 13.8, 5.2 Hz, H-17), 3.32 (1H, d, J = 4.0 Hz, H-3), 3.29 (1H, d, J = 4.0 Hz, H-2), 2.85 (1H, dd, J = 6.5, 4.0 Hz, H-11), 2.21 (1H, dd, J = 13.6, 6.5 Hz, H-12a), 2.16 (1H, ddd, J = 13.8, 12.5, 3.2 Hz, H-16a), 1.95 (1H, ddd, J = 12.5, 5.2, 1.0 Hz, H-16b), 1.65 (3H, s, H-30), 1.62 (1H, s, H-7), 1.54 (3H, s, H-19), 1.53 (1H, brs, OH-15), 1.36 (1H, dd, J = 13.6, 4.0 Hz, H-12b), 1.31 (3H, s, H-29), 0.8 (3H, s, H-18). 13C NMR (150 MHz, CDCl3) 205.1 (C-1), 151.9 (C-6), 142.8 (C-23), 139.5 (C-21), 124.5 (C-20), 115.6 (C-5), 111.2 (C-22), 80.1 (C-28), 79.8 (C-15), 63.1 (C-3), 62.0 (C-9), 59.3 (C-11), 57.1 (C-2), 51.3 (C-14), 48.9 (C-10), 45.9 (C-4), 45.5 (C-17), 43.7 (C-13), 40.9 (C-12), 38.7 (C-16), 29.5 (C-7), 23.7 (C-8), 23.1 (C-19), 23.08 (C-29), 19.5 (C-18), 19.1 (C-30).

3.4. Calculation of ECD Spectra

Preliminary conformational analyses of all compounds, except compound 4, were performed by CONFLEX9 with the MMFF94 force field. Spaltan20 was used for the preliminary conformational analysis of compound 4. The obtained conformers were further optimized in MeOH by the density functional theory (DFT) method, with the B3LYP functional and 6–31(d) basis set. The ECD spectrum was calculated via the time-dependent DFT (TDDFT) method, using the CAM-B3LYP functional and TZVP basis set. The calculation was performed using the conformers within 2 kcal/mol predicted in MeOH. The solvent effect was introduced by the conductor-like polarizable continuum model (CPCM). The DFT optimization and TDDFT-ECD calculation were accomplished by Gaussian16 (Gaussian, Inc., Wallingford, CT, USA). The calculated spectrum was displayed using GaussView 6.1, with the peak half-width at half-height being 0.333 eV. The Boltzmann-averaged spectrum at 298.15 K was calculated using Excel 2016 (Microsoft Co., Redmond, WA, USA). The calculations were re-optimized according to the literature [41].

3.5. Antiproliferative Activity Assay

A549, KB, MDA-MB-231 and MCF-7 were obtained from the Lineberger comprehensive Cancer Center (UNC-CH, NC) or from ATCC (Manassas, VA, USA). KB-VIN was a generous gift of Professor Y.-C. Cheng (Yale University, New Haven, CT, USA). We confirmed our KB and KB-VIN are identical to AV-3 (ATCC number, CCL-21) as a HeLa (cervical carcinoma) derivative by short tandem repeat (STR) profiling. The antiproliferative activity was examined via an SRB assay, as described previously [42]. Briefly, freshly trypsinized cell suspensions were seeded in 96-well microtiter plates at densities of 4000–11,000 cells per well for each compound. The attached cells were fixed in 10% trichloroacetic acid and stained with 0.04% SRB after culturing for 72 h. The absorbance at 515 nm was measured using a microplate reader (Spark 10M, Tecan, Zurich, Switzerland) with SparkControl software version 2.3 (Tecan) after solubilizing the bound dye with 10 mM Tris base. The mean IC50 was determined as the average from at least three independent experiments of duplication for an assay and similar determinations.

4. Conclusions

The phytochemical investigation of a 50% MeOH/CH2Cl2 extract from the leaves of a tropical rainforest plant, T. rubescens, led to the isolation of 14 vilasinin-class limonoids, including the new rubescins K–R (18) and known rubescins E, F, and H–J (913), as well as TS3 (14), together with three sesquiterpenes and two sterols. An extensive analysis of isolated compounds revealed that rubescins K–M (13) were unusually chlorinated limonoids and rubescin N (4) was the first 2,3-epoxylated vilasinin-class limonoid. The natural occurrence of chlorinated limonoids was further confirmed using chemical methods and LC/MS analysis.
The isolated vilasinin-type limonoids 1, 5, 6, 8, 9, 11, 13, and 14 were evaluated for their growth-inhibitory effects against five human tumor cell lines, including a multidrug-resistant cell line, KB-VIN. All the tested limonoids, except for compound 8, exhibited significant activity, with IC50 values of 0.54–8.46 μM against all tested tumor cell lines, including KB-VIN. Compound 14 showed the highest inhibitory activity, while chlorinated limonoid 1 demonstrated slightly selective inhibition against the MCF-7 cell line. The preliminary structure–activity relationship study indicated that the tetrahydrofuran ring formed by C4–6 and C-28, which is characteristic of the vilasinin-class limonoid, might be important for this activity.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules29030651/s1, Figure S1. NCI-60 Human tumor cell line assay data for the crude organic extract of T. rubescens (N047159). Figure S2. 1H NMR spectrum of compound 1 in CDCl3 at 600 MHz. Figure S3. 13C NMR spectrum of compound 1 in CDCl3 at 150 MHz. Figure S4. HMQC spectrum of compound 1 in CDCl3 at 600 MHz. Figure S5. HMBC spectrum of compound 1 in CDCl3 at 600 MHz. Figure S6. COSY spectrum of compound 1 in CDCl3 at 600 MHz. Figure S7. NOESY spectrum of compound 1 in CDCl3 at 600 MHz. Figure S8. UV spectrum of compound 1 in MeOH. Figure S9. IR spectrum of compound 1. Figure S10. 1H NMR spectrum of compound 2 in CDCl3 at 600 MHz. Figure S11. 13C NMR spectrum of compound 2 in CDCl3 at 150 MHz. Figure S12. HMQC spectrum of compound 2 in CDCl3 at 600 MHz. Figure S13. HMBC spectrum of compound 2 in CDCl3 at 600 MHz. Figure S14. COSY spectrum of compound 2 in CDCl3 at 600 MHz. Figure S15. NOESY spectrum of compound 2 in CDCl3 at 600 MHz. Figure S16. 1H NMR spectrum of compound 3 in CDCl3 at 600 MHz. Figure S17. 13C NMR spectrum of compound 3 in CDCl3 at 150 MHz. Figure S18. HMQC spectrum of compound 3 in CDCl3 at 600 MHz. Figure S19. HMBC spectrum of compound 3 in CDCl3 at 600 MHz. Figure S20. COSY spectrum of compound 3 in CDCl3 at 600 MHz. Figure S21. NOESY spectrum of compound 3 in CDCl3 at 600 MHz. Figure S22. 1H NMR spectrum of compound 4 in CDCl3 at 600 MHz. Figure S23. 13C NMR spectrum of compound 4 in CDCl3 at 150 MHz. Figure S24. HMQC spectrum of compound 4 in CDCl3 at 600 MHz. Figure S25. HMBC spectrum of compound 4 in CDCl3 at 600 MHz. Figure S26. COSY spectrum of compound 4 in CDCl3 at 600 MHz. Figure S27. NOESY spectrum of compound 4 in CDCl3 at 600 MHz. Figure S28. UV spectrum of compound 4 in MeOH. Figure S29. IR spectrum of compound 4. Figure S30. 1H NMR spectrum of compound 5 in CDCl3 at 600 MHz. Figure S31. 13C NMR spectrum of compound 5 in CDCl3 at 150 MHz. Figure S32. HMQC spectrum of compound 5 in CDCl3 at 600 MHz. Figure S33. HMBC spectrum of compound 5 in CDCl3 at 600 MHz. Figure S34. COSY spectrum of compound 5 in CDCl3 at 600 MHz. Figure S35. NOESY spectrum of compound 5 in CDCl3 at 600 MHz. Figure S36. UV spectrum of compound 5 in MeOH. Figure S37. IR spectrum of compound 5. Figure S38. 1H NMR spectrum of compound 6 in CDCl3 at 600 MHz. Figure S39.13C NMR spectrum of compound 6 in CDCl3 at 150 MHz. Figure S40. HMQC spectrum of compound 6 in CDCl3 at 600 MHz. Figure S41. HMBC spectrum of compound 6 in CDCl3 at 600 MHz. Figure S42. COSY spectrum of compound 6 in CDCl3 at 600 MHz. Figure S43. NOESY spectrum of compound 6 in CDCl3 at 600 MHz. Figure S44. UV spectrum of compound 6 in MeOH. Figure S45. IR spectrum of compound 6. Figure S46. 1H NMR spectrum of compound 7 in CDCl3 at 600 MHz. Figure S47. 13C NMR spectrum of compound 7 in CDCl3 at 150 MHz. Figure S48. HMQC spectrum of compound 7 in CDCl3 at 600 MHz. Figure S49. HMBC spectrum of compound 7 in CDCl3 at 600 MHz. Figure S50. COSY spectrum of compound 7 in CDCl3 at 600 MHz. Figure S51. NOESY spectrum of compound 7 in CDCl3 at 600 MHz. Figure S52. UV spectrum of compound 7 in MeOH. Figure S53. IR spectrum of compound 7. Figure S54. 1H NMR spectrum of compound 8 in CDCl3 at 600 MHz. Figure S55.13C NMR spectrum of compound 8 in CDCl3 at 150 MHz. Figure S56. HMQC spectrum of compound 8 in CDCl3 at 600 MHz. Figure S57. HMBC spectrum of compound 8 in CDCl3 at 600 MHz. Figure S58. COSY spectrum of compound 8 in CDCl3 at 600 MHz. Figure S59. NOESY spectrum of compound 8 in CDCl3 at 600 MHz. Figure S60. UV spectrum of compound 8 in MeOH. Figure S61. IR spectrum of compound 8. Figure S62. 1H NMR spectrum of compound 17 in CDCl3 at 600 MHz. Figure S63.13C NMR spectrum of compound 17 in CDCl3 at 150 MHz. Figure S64. LC/MS analysis of compound 1. Figure S65. LC/MS analysis of fraction 7.2. Figure S66. Structures of TS1 and TS2.

Author Contributions

Conceptualization, K.N.-G., D.J.N., B.R.O. and K.-H.L.; methodology, K.N.-G., K.M., Y.S., D.J.N. and K.-H.L.; validation, K.N.-G., D.J.N., B.R.O., K.M. and Y.S.; investigation, S.A., S.F. and Y.S.; resources, D.J.N., B.R.O., K.-H.L. and K.N.-G.; data curation, S.A., Y.S., S.F. and K.N.-G.; writing—original draft preparation, K.N.-G. and S.A.; writing—review and editing, K.N.-G.; supervision, K.N.-G.; project administration, K.N.-G.; funding acquisition, K.N.-G., Y.S. and S.A. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by a Grant-in-Aid from JSPS KAKENHI, Japan (grant number 18H02583), awarded to K.N.-G. This work was also partially supported by a grant from the Shibuya Science Culture and Sports Foundation, awarded to Y.S. and K.N.-G. This study was supported by a JST SPRING grant (grant number JPMJSP2135) awarded to S.A.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

We wish to thank Susan. L. Morris-Natschke (UNC-CH) for critical comments, suggestions, and editing on the manuscript. We thank Yusuke Masuo (Kanazawa University) for his assistance in performing LC/MS to provide the mass spectra of compound 1.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Luo, J.; Sun, Y.; Li, Q.; Kong, L. Research progress of meliaceous limonoids from 2011 to 2021. Nat. Prod. Rep. 2022, 39, 1325–1365. [Google Scholar] [CrossRef]
  2. Gouvea, C.F.; Dornelas, M.C.; Rodriguez, A.P.M. Floral development in the tribe Cedreleae (Meliaceae, sub-family Swietenioideae): Cedrela and Toona. Ann. Bot. 2008, 101, 39–48. [Google Scholar] [CrossRef] [PubMed]
  3. Passos, M.S.; Nogueira, T.S.R.; Azevedo, O.d.A.; Vieira, M.G.C.; Terra, W.d.S.; Braz-Filho, R.; Vieira, I.J.C. Limonoids from the genus Trichilia and biological activities: Review. Phytochem. Rev. 2021, 20, 1055–1086. [Google Scholar] [CrossRef]
  4. Hilmayanti, E.; Nurlelasari; Supratman, U.; Kabayama, K.; Shinoyama, A.; Kukase, K. Limonoids with anti-inflammatory activity: A review. Phytochemistry 2022, 204, 113469. [Google Scholar] [CrossRef] [PubMed]
  5. Tsamo, A.T.; Mkounga, P.; Njayou, N.; Manautou, J.; Hultin, P.G.; Nkengfack, A.E. Rubescins A, B and C: New Havanensin Type Limonoids from Root Bark of Trichilia rubescens (Meliaceae). Chem. Pharm. Bull. 2013, 61, 1178–1183. [Google Scholar] [CrossRef] [PubMed]
  6. Armelle, T.T.; Pamela, N.K.; Pierre, M.; Muller, I.B.; Marat, K.; Sass, G.; Nkengfack, A.E. Antiplasmodial Limonoids from Trichilia rubescens (Meliaceae). Med. Chem. 2016, 12, 655–661. [Google Scholar] [CrossRef] [PubMed]
  7. Tsamo, A.T.; Pagna, J.I.M.; Nangmo, P.K.; Mkounga, P.; Laatsch, H.; Nkengfack, A.E. Rubescins F–H, new vilasinin-type limonoids from the leaves of Trichilia rubescens (Meliaceae). Z. Naturforsch. C J. Biosci. 2019, 74, 175–182. [Google Scholar] [CrossRef]
  8. Tsamo, A.T.; Melong, R.; Mkounga, P.; Nkengfack, A.E. Rubescins I and J, further limonoid derivatives from the stem bark of Trichilia rubescens (Meliaceae). Nat. Prod. Res. 2019, 33, 196–203. [Google Scholar] [CrossRef]
  9. De Carvalho, A.C.V.; Ndi, C.P.; Tsopmo, A.; Tane, P.; Ayafor, J.; Connolly, J.D.; Teem, J.L. A Novel Natural Product Compound Enhances cAMP-Regulated Chloride Conductance of Cells Expressing CFTR∆F508. Mol. Med. 2002, 8, 75–87. [Google Scholar] [CrossRef]
  10. Krief, S.; Martin, M.T.; Greller, P.; Kasenene, J.; Sevenet, T. Novel Antimalarial Compounds Isolated in a Survey of Self-Medicative Behavior of Wild Chimpanzees in Uganda. Antimicrob. Agents Chemother. 2004, 48, 3196–3199. [Google Scholar] [CrossRef]
  11. Nakatani, M.; Iwashita, T.; Mizukawa, K.; Hase, T. Trichilinin, a New Hexacyclic Limonoid from Trichilia roka. Heterocycles 1987, 26, 43–46. [Google Scholar] [CrossRef]
  12. Kowa, T.K.; Jansen, O.; Ledoux, A.; Mamede, L.; Wabo, H.K.; Tchinda, A.T.; Gregory, G.J.; Frederich, M. Bioassay-guided isolation of vilasinin–type limonoids and phenyl alkene from the leaves of Trichilia gilgiana and their antiplasmodial activities. Nat. Prod. Res. 2022, 36, 5039–5047. [Google Scholar] [CrossRef] [PubMed]
  13. El Sayed, K.A.; Hamann, M.T. A New Norcembranoid Dimer from the Red Sea Soft Coral Sinularia gardineri. J. Nat. Prod. 1996, 59, 687–689. [Google Scholar] [CrossRef] [PubMed]
  14. Oshima, Y.; Iwakawa, T.; Hikino, H. Alismol and alismoxide, sesquiterpenoids of Alisma rhizomes. Phytochemistry 1983, 22, 183–185. [Google Scholar] [CrossRef]
  15. Zhang, H.J.; Tan, G.T.; Santarsiero, B.D.; Mesecar, A.D.; Hung, N.V.; Cuong, N.M.; Soejarto, D.D.; Pezzuto, J.M.; Fong, H.S. New Sesquiterpenes from Litsea verticillate. J. Nat. Prod. 2003, 66, 609–615. [Google Scholar] [CrossRef] [PubMed]
  16. Nishioka, I.; Ikekawa, N.; Yagi, A.; Kawasaki, T.; Tsukamoto, T. Studies on the Plant Sterols and Triterpenes. II. Separation of Stigmasterol, β-Sitosterol and Campesterol, and about So-called “γ-Sitosterol.”. Chem. Pharm. Bull. 1965, 13, 379–384. [Google Scholar] [CrossRef]
  17. Li, S.; Li, Y.; Xu, R.; Kong, L.Y.; Luo, J. New meliacarpin-type (C-seco) and C-ring intact limonoids from the fruits of Melia toosendan. Fitoterapia 2020, 144, 104605. [Google Scholar] [CrossRef]
  18. Liu, H.B.; Zhang, C.R.; Dong, S.H.; Dong, L.; Wu, Y.; Yue, J.M. Limonoids and Triterpenoids from the Seeds of Melia azedarach. Chem. Pharm. Bull. 2011, 59, 1003–1007. [Google Scholar] [CrossRef]
  19. Nakai, H.; Shiro, M.; Ochi, M. Ohchinolide A, a new limonoid from Melia azedarach L. var. japonica Makino. Acta Cryst. 1980, 36, 1698–1700. [Google Scholar] [CrossRef]
  20. Akihisa, T.; Pan, X.; Nakamura, Y.; Kikuchi, T.; Takahashi, N.; Matsumoto, M.; Ogihara, E.; Fukatsu, M.; Koike, K.; Tokuda, H. Limonoids from the fruits of Melia azedarach and their cytotoxic activities. Phytochemistry 2013, 89, 59–70. [Google Scholar] [CrossRef]
  21. Zhu, G.Y.; Bai, L.P.; Liu, L.; Jiang, Z.H. Limonoids from the fruits of Melia toosendan and their NF-κB modulating activities. Phytochemistry 2014, 107, 175–181. [Google Scholar] [CrossRef] [PubMed]
  22. Fukuyama, Y.; Nakaoka, M.; Yamamoto, T.; Takahshi, H.; Minami, H. Degraded and Oxetane-Bearing Limonoids from the Roots of Melia azedarach. Chem. Pharm. Bull. 2006, 54, 1219–1222. [Google Scholar] [CrossRef] [PubMed]
  23. Dong, S.H.; Zhang, C.R.; He, X.F.; Liu, H.B.; Wu, Y.; Yue, J.M. Mesendanins A−J, Limonoids from the Leaves and Twigs of Melia toosendan. J. Nat. Prod. 2010, 73, 1344–1349. [Google Scholar] [CrossRef]
  24. Zhang, Y.; Tang, C.P.; Ke, C.Q.; Li, X.Q.; Xie, H.; Ye, Y. Limonoids from the fruits of Melia toosendan. Phytochemistry 2012, 73, 106–113. [Google Scholar] [CrossRef]
  25. Zeng, J.; Zhang, J. Chlorinated Natural Products and Related Halogenases. Isr. J. Chem. 2019, 59, 387–402. [Google Scholar] [CrossRef]
  26. Adak, S.; Moore, B.S. Cryptic halogenation reactions in natural product biosynthesis. Nat. Prod. Rep. 2021, 38, 1760–1774. [Google Scholar] [CrossRef]
  27. Gribble, G.W. A recent survey of naturally occurring organohalogen compounds. Environ. Chem. 2015, 12, 396–405. [Google Scholar] [CrossRef]
  28. Wang, L.; Zhou, X.; Fredimoses, M.; Liao, S.; Liu, Y. Naturally occurring organoiodines. RSC Adv. 2014, 4, 57350–57376. [Google Scholar] [CrossRef]
  29. Bidleman, T.F.; Andersson, A.; Jantunen, L.M.; Kucklick, J.R.; Kylin, H.; Letcher, R.J.; Tysklind, M.; Wong, F. A review of halogenated natural products in Arctic, Subarctic and Nordic ecosystems. Emerg. Contam. 2019, 5, 89–115. [Google Scholar] [CrossRef]
  30. Wagner, C.; Omari, M.E.; Konig, G.M. Biohalogenation: Nature’s Way to Synthesize Halogenated Metabolites. J. Nat. Prod. 2009, 72, 540–553. [Google Scholar] [CrossRef]
  31. Crowe, C.; Molyneux, S.; Sharma, S.V.; Zhang, Y.; Gkotsi, D.S.; Connaris, H.; Goss, R.J.M. Halogenases: A palette of emerging opportunities for synthetic biology–synthetic chemistry and C–H functionalization. Chem. Soc. Rev. 2021, 50, 9443–9481. [Google Scholar] [CrossRef]
  32. Engvild, K.C. Chlorine-containing natural compounds in higher plants. Phytochemistry 1986, 25, 781–791. [Google Scholar] [CrossRef]
  33. Fehlberg, I.; Riberio, P.R.; dos Santos, I.B.F.; dos Sasntos, I.I.P.; Guedes, M.L.S.; Ferraz, C.G.; Cruz, F.G. Two new sesquiterpenoids and one new p-coumaroyl-triterpenoid derivative from Myrcia guianensis. Phytochem. Lett. 2022, 48, 5–10. [Google Scholar] [CrossRef]
  34. Ang, S.; Liu, C.; Hong, P.; Yang, L.; Hu, G.; Zheng, X.; Jin, J.; Wu, R.; Wong, W.L.; Zhang, K.; et al. Hirsutinolide-type sesquiterpenoids with anti-prostate cancer activity from Cyanthillium cinereum. Phytochemistry 2023, 216, 113887. [Google Scholar] [CrossRef]
  35. Li, H.; Liu, L.; Liu, G.; Li, J.; Aisa, H. A Chlorine-containing guaianolide sesquiterpenoids from Achillea millefolium L. with inhibitory effects against LPS-induced NO release in BV-2 microglial cells. Phytochemistry 2023, 207, 113567. [Google Scholar] [CrossRef]
  36. Yadav, P.A.; Kumar, C.P.; Siva, B.; Babu, K.S.; Allanki, A.D.; Sijwali, P.S.; Jain, N.; Rao, A.V. Synthesis and evaluation of anti-plasmodial and cytotoxic activities of epoxyazadiradione derivatives. Eur. J. Med. Chem. 2017, 134, 242–257. [Google Scholar] [CrossRef]
  37. Loughlin, W.A.; Jenkins, I.D.; Henderson, L.C.; Campitelli, M.R.; Healy, P. Total Synthesis of (±)-Hyphodermins A and D. J. Org. Chem. 2008, 73, 3435–3440. [Google Scholar] [CrossRef] [PubMed]
  38. Lange, N.; Tontsa, A.T.; Wegscheid, C.; Mkounga, P.; Nkengfack, A.E.; Loscher, C.; Sass, G.; Tiegs, G. The Limonoids TS3 and Rubescin E Induce Apoptosis in Human Hepatoma Cell Lines and Interfere with NF-κB Signaling. PLoS ONE 2016, 12, 11. [Google Scholar] [CrossRef] [PubMed]
  39. Mohamad, K.; Hirasawa, Y.; Litaudon, M.; Awang, K.; Hadi, A.A.H.; Takeya, K.; Ekasari, W.; Widyawaruyanti, A.; Zaini, N.C.; Morita, H. Ceramicines B–D, new antiplasmodial limonoids from Chisocheton ceramicus. Bioorg. Med. Chem. 2009, 17, 727–730. [Google Scholar] [CrossRef] [PubMed]
  40. McCloud, T.G. High Throughput Extraction of Plant, Marine and Fungal Specimens for Preservation of Biologically Active Molecules. Molecules 2010, 15, 4526–4563. [Google Scholar] [CrossRef] [PubMed]
  41. Pascitelli, G.; Bruhn, T. Good Computational Practice in the Assignment of Absolute Configurations by TDDFT Calculations of ECD Spectra. Chirality 2016, 28, 466–474. [Google Scholar] [CrossRef] [PubMed]
  42. Hirazawa, S.; Saito, Y.; Sagano, M.; Goto, M.; Nakagawa-Goto, K. Chemical Space Expansion of Flavonoids: Induction of Mitotic Inhibition by Replacing Ring B with a 10π-Electron System, Benzo[b]thiophene. J. Nat. Prod. 2022, 85, 136–147. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Structure of vilasinin, a ring-intact limonoid.
Figure 1. Structure of vilasinin, a ring-intact limonoid.
Molecules 29 00651 g001
Figure 2. New vilasinin-class limonoids 18 isolated from T. rubescens.
Figure 2. New vilasinin-class limonoids 18 isolated from T. rubescens.
Molecules 29 00651 g002
Figure 3. Known vilasinin-class limonoids 914 isolated from T. rubescens in this work.
Figure 3. Known vilasinin-class limonoids 914 isolated from T. rubescens in this work.
Molecules 29 00651 g003
Figure 4. Selected 1H–1H COSY and HMBC correlations of new limonoids 18.
Figure 4. Selected 1H–1H COSY and HMBC correlations of new limonoids 18.
Molecules 29 00651 g004
Figure 5. Selected NOESY correlations (red dashed arrows) of compounds 18.
Figure 5. Selected NOESY correlations (red dashed arrows) of compounds 18.
Molecules 29 00651 g005
Figure 6. Experimental and calculated ECD spectra of compounds 1 and 48.
Figure 6. Experimental and calculated ECD spectra of compounds 1 and 48.
Molecules 29 00651 g006
Figure 7. Possible biosynthetic pathway of chlorinated vilasinin-class limonoids.
Figure 7. Possible biosynthetic pathway of chlorinated vilasinin-class limonoids.
Molecules 29 00651 g007
Figure 8. Investigation of the possibility of an artifact using a model substrate 11.
Figure 8. Investigation of the possibility of an artifact using a model substrate 11.
Molecules 29 00651 g008
Table 1. 1H NMR spectroscopic data (600 MHz, CDCl3) of chlorinated limonoids 13.
Table 1. 1H NMR spectroscopic data (600 MHz, CDCl3) of chlorinated limonoids 13.
Position123
δH (J in Hz)δH (J in Hz)δH (J in Hz)
24.74, d (10.1)4.72, d (10.3)4.70, d (10.1)
33.87, d (10.1)3.77, d (10.3)3.84, d (10.1)
52.54, d (11.9)2.38, d (11.9)2.43, d (11.9)
64.59, dd (11.9, 3.8)4.51, dd (11.9, 3.8)4.54, dd (11.9, 3.9)
75.66, d (3.8)5.47, d (3.8)5.48, d (3.8)
114.68, dd (6.9, 1.4)4.65, dd (6.9, 1.4)4.65, dd (6.8, 1.3)
12a1.97, dd (13.7, 6.9)1.98, dd (13.3, 6.9)2.03, dd (13.5, 6.8)
12b1.84, dd (13.7, 1.4)1.83, overlap1.87, d (13.5)
153.53, brs a3.46, brs a3.45, brs a
16a2.07, ddd (13.3, 6.5, 1.0)2.11, ddd (13.4, 6.5, 0.7)2.12, ddd (13.4, 6.4, 1.0)
16b1.42, dd (13.3, 11.3)1.54, overlap1.59, dd (13.4, 11.3)
172.47, dd (11.3, 6.5)2.51, dd (11.3, 6.5)2.53, dd (11.3, 6.4)
180.61, s0.62, s0.77, s
191.54, s1.50, s1.51, s
216.88, dd (1.6, 0.9)7.03, brs7.01, dd (1.6, 0.9)
225.87, dd (1.6, 0.9)6.07, brs6.07, dd (1.6, 0.9)
237.24, t (1.6)7.35, t (1.7)7.31, t (1.6)
28a3.85, d (7.3)3.91, d (7.9)3.92, d (7.8)
28b3.48, d (7.3)3.52, d (7.9)3.54, d (7.8)
291.40, s1.39, s1.38, s
301.26, s1.20, s 1.22, s
OH-32.68, brs a2.61, brs a2.58, brs a
Ester moiety at C-7
positionδH (J in Hz)positionδH (J in Hz)positionδH (J in Hz)
2′ 2′ 2′6.37, d (15.9)
3′/7′7.92, dd (8.3, 1.3)3′6.72, dd (7.2, 1.4)3′7.79, d (15.9)
4′/6′7.47, dd (8.3, 8.3)4′1.83, d (7.2)4′-
5′7.62, m5′1.86, brs5′/9′7.53, dd (7.8, 2.2)
6′/8′7.44, m
7′7.43, m
a Broad singlet.
Table 2. 13C NMR spectroscopic data (150 MHz, CDCl3) of chlorinated limonoids 13.
Table 2. 13C NMR spectroscopic data (150 MHz, CDCl3) of chlorinated limonoids 13.
Position123
δCδCδC
1199.4199.3199.5
268.868.768.7
379.479.579.2
445.445.345.2
548.248.147.8
671.371.371.3
774.974.174.1
845.145.245.0
964.964.964.9
1050.150.150.0
1160.260.260.2
1235.135.335.2
1341.141.141.2
1468.468.368.4
1555.655.655.5
1631.131.131.1
1738.638.838.7
1821.721.418.1
1918.118.221.2
20122.7123.0122.9
21139.3139.4139.4
22110.8110.9110.9
23142.8142.9143.0
2883.082.983.0
2914.314.414.3
3023.022.822.9
Ester moiety at C-7
positionδCpositionδCpositionδC
1′165.91′167.21′165.9
2′129.82′128.72′116.7
3′/7′128.73′138.53′147.1
4′/6′129.64′14.64′133.9
5′133.65′12.75′/9′128.3
6′/8′129.1
7′130.9
Table 3. 1H NMR spectroscopic data (600MHz, CDCl3) of compounds 48.
Table 3. 1H NMR spectroscopic data (600MHz, CDCl3) of compounds 48.
Position45678
δH (J in Hz)δH (J in Hz)δH (J in Hz)δH (J in Hz)δH (J in Hz)
23.15, d (2.8)5.99, d (9.6)5.87, d (9.6)5.85, d (9.8)5.90, d (10.3)
33.57, d (2.8)7.10, d (9.6)6.94, d (9.6)6.68, d (9.8)6.64, d (10.3)
53.12, d (12.6)3.05, d (12.7)2.66, d (12.4) 3.67, s
64.00, dd (12.6, 4.1)4.49, dd (12.7, 4.1)4.34, dd (12.4, 3.1)
75.02, d (4.1)5.68, d (4.1)4.09, d (3.1)1.66, s2.00, s
9 2.44, d (3.7)
113.18, dd (6.5, 4.7)3.98, dd (6.5, 1.4)6.23, dd (8.0, 3.7)2.69, dd (6.3, 4.3)3.04, dd (6.5, 3.6)
12a2.21, dd (13.8, 6.5)1.90, dd (13.6, 6.5)2.50, dd (14.6, 8.0)2.60, dd (13.5, 6.3)2.27, dd (13.8, 6.5)
12b1.31, overlap1.85, d (13.6)1.64, dd (14.6, 3.7)1.47, dd (13.5, 4.3)1.51, dd (13.8, 3.6)
154.66, brs a3.50, brs a5.67, brs a3.90, brs a4.00, brs a
16a2.08, ddd (13.9, 13.0, 2.9)2.07, dd (13.2, 6.5)2.53, ddd (15.6,
11.3, 1.7)
2.17, ddd (13.0,
11.6, 3.6)
2.20, ddd (13.8,
12.6, 3.4)
16b1.85, ddd (13.0,
5.3, 1.0)
1.41, dd (13.2, 11.3)2.45, ddd (15.6,
7.6, 3.4)
1.95/1.93, dd (11.6, 5.6)1.95, ddd (12.6, 4.8, 0.9)
173.12, dd (13.9, 5.3)2.47, dd (11.3, 6.5)2.89, dd (11.3, 7.6)3.34/3.32, dd (13.0, 5.6)3.43, dd (13.8, 4.8)
180.76, s0.64, s0.82, s0.66/0.63, s0.67, s
191.48, s1.46, s1.35, s1.64/1.63, s1.67, s
217.17, dd (1.6, 0.8)6.84, brs a7.25, brs a 7.21, dd (1.7, 0.8)
226.20, dd (1.6, 0.8)5.87, brs a6.30, brs a6.85, s6.23, dd (1.7, 0.8)
237.37, t (1.6)7.22, t (1.7)7.37, t (1.7)6.12, brs7.21, t (1.7)
28a3.96, d (7.7)3.70, d (7.4)3.79, d (7.2)4.29, d (9.0)4.20, dd (10.2, 7.1)
28b3.92, d (7.7)3.54, d (7.4)3.62, d (7.2)4.07/4.05, d (9.0)3.45, dd (10.2, 7.1)
291.26, s1.35, s1.36, s1.36, s1.20, s
301.31, s1.27, s1.46, s1.68/1.67, s1.67, s
OH-7 2.09, s
OH-151.42, d (2.9) 1.55, overlap1.56, d (2.8)
OH-28 1.64, d (7.1)
2′ 2.04, s
3′/7′ 7.94, dd (7.9, 1.4)
4′/6′ 7.46, dd (7.9, 7.9)
5′ 7.59, m
a Broad singlet.
Table 4. 13C NMR spectroscopic data (150MHz, CDCl3) of compounds 48.
Table 4. 13C NMR spectroscopic data (150MHz, CDCl3) of compounds 48.
Position4567 a8
δCδCδCδCδC
1202.4200.3202.1197.1196.7
251.9131.3130.0128.54/128.53127.6
360.4151.7150.8152.4/152.3155.4
439.042.841.944.842.1
551.850.848.0118.255.4
671.871.373.8149.7207.1
780.075.575.429.144.4
847.944.845.922.729.3
961.864.639.460.561.3
1047.847.647.348.249.3
1159.460.471.058.758.1
1237.235.344.240.9/40.739.9
1345.641.445.543.8/43.743.9
1496.168.3158.451.4, 51.355.2
1576.155.2121.679.080.3
1637.131.134.337.5/37.438.6
1743.038.752.345.0/44.946.0
1819.921.822.519.422.3
1915.917.416.321.921.0
20123.9122.8124.1138.8/138.9124.1
21139.5139.3139.8171.3/171.0139.6
22111.1110.9110.9144.5/144.9110.9
23142.9142.7142.896.7/96.3143.1
2879.479.780.181.470.3
2916.821.120.128.117.4
3021.622.827.819.218.9
1′ 166.1169.7
2′ 128.621.8
3′/7′ 129.9
4′/6′ 129.8
5′ 133.4
a 1:1 Diastereomixture.
Table 5. Antiproliferative activity of the selected limonoids.
Table 5. Antiproliferative activity of the selected limonoids.
Cell Lines a (IC50 μM)
CompoundsA549MDA-MB-231MCF-7KBKB-VIN
12.535.721.343.384.02
53.175.452.854.482.09
63.3917.64.566.838.07
837.4>40>40>40>40
92.145.880.963.684.35
111.6020.98.466.838.09
131.426.768.074.383.28
140.562.060.750.570.54
PXL b (nM)2.5712.412.716.762939.33
a A549: Lung adenocarcinoma, MDA-MB-231: triple-negative breast cancer (ER-/PR-/HER2-), MCF-7: HER2-negative, KB: originally isolated from an epidermoid carcinoma of the nasopharynx (contaminated HeLa cell), KB-VIN: multidrug-resistant KB subline. b Paclitaxel.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Amuti, S.; Saito, Y.; Fukuyoshi, S.; Miyake, K.; Newman, D.J.; O’Keefe, B.R.; Lee, K.-H.; Nakagawa-Goto, K. Unusual Vilasinin-Class Limonoids from Trichilia rubescens. Molecules 2024, 29, 651. https://doi.org/10.3390/molecules29030651

AMA Style

Amuti S, Saito Y, Fukuyoshi S, Miyake K, Newman DJ, O’Keefe BR, Lee K-H, Nakagawa-Goto K. Unusual Vilasinin-Class Limonoids from Trichilia rubescens. Molecules. 2024; 29(3):651. https://doi.org/10.3390/molecules29030651

Chicago/Turabian Style

Amuti, Saidanxia, Yohei Saito, Shuichi Fukuyoshi, Katsunori Miyake, David J. Newman, Barry R. O’Keefe, Kuo-Hsiung Lee, and Kyoko Nakagawa-Goto. 2024. "Unusual Vilasinin-Class Limonoids from Trichilia rubescens" Molecules 29, no. 3: 651. https://doi.org/10.3390/molecules29030651

Article Metrics

Back to TopTop