Next Article in Journal
Effect of Ylang-Ylang (Cananga odorata Hook. F. & Thomson) Essential Oil on Acute Inflammatory Response In Vitro and In Vivo
Previous Article in Journal
Nanocomposites Based on Antiferroelectric Liquid Crystal (S)-MHPOBC Doping with Au Nanoparticles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

New Terpenoids from Potentilla freyniana Bornm. and Their Cytotoxic Activities

1
TCM and Ethnomedicine Innovation and Development International Laboratory, School of Pharmacy, Hunan University of Chinese Medicine, Changsha 410208, China
2
Hunan Provincial Key Laboratory of Dong Medicine, Hunan University of Medicine, Huaihua 418000, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2022, 27(12), 3665; https://doi.org/10.3390/molecules27123665
Submission received: 17 May 2022 / Revised: 2 June 2022 / Accepted: 5 June 2022 / Published: 7 June 2022
(This article belongs to the Section Natural Products Chemistry)

Abstract

:
Two new A-ring contracted triterpenoids, madengaisu A and madengaisu B, and one undescribed ent-kaurane diterpenoid, madengaisu C, along with 20 known compounds were isolated from the roots of Potentilla freyniana Bornm. The structures were elucidated using extensive spectroscopic techniques, including 1D and 2D-NMR, HR-ESI-MS, ECD spectra, IR, and UV analysis. Moreover, all isolated constituents were evaluated for their anti-proliferative activity against RA-FLS cells and cytotoxic activities against the human cancer cell lines Hep-G2, HCT-116, BGC-823, and MCF-7. Ursolic acid and pomolic acid displayed moderate inhibitory activity in RA-FLS cells with IC50 values of 24.63 ± 1.96 and 25.12 ± 1.97 μM, respectively. Hyptadienic acid and 2α,3β-dihydroxyolean-12-en-28-oic acid 28-O-β-d-glucopyranoside exhibited good cytotoxicity against Hep-G2 cells with IC50 values of 25.16 ± 2.55 and 17.66 ± 1.82 μM, respectively. In addition, 2α,3β-dihydroxyolean-13(18)-en-28-oic acid and alphitolic acid were observed to inhibit HCT-116 cells (13.25 ± 1.65 and 21.62 ± 0.33 μM, respectively), while madengaisu B and 2α,3β-dihydroxyolean-13(18)-en-28-oic acid showed cytotoxic activities against BGC-823 cells with IC50 values of 24.76 ± 0.94 and 26.83 ± 2.52 μM, respectively, which demonstrated that triterpenes from P. freyniana may serve as therapeutic agents for RA and cancer treatment.

1. Introduction

Potentilla L. is a major genus belonging to the family Rosaceae with more than 200 species in the world, and many of them are widely used as ethnomedicine. Previous phytochemical research has shown that the genus Potentilla L. contains amounts of bioactive compounds, such as triterpenoids, flavonoids, and tannins, etc. Pharmacological studies also reported that Potentilla species possess significant hypoglycemic, anti-inflammatory, anti-tumor, antioxidative, and antibacterial activities [1].
Potentilla freyniana Bornm. is a perennial herb and is widely distributed in China, Russia, Japan, and Korea. It is also called “Ma Deng Ai” or “San Zhang ye” in the Dong ethnomedicine of China. The commonly used medicinal parts of the herb include the roots and whole grass, which have the effect of clearing heat and detoxifying, dispersing blood stasis, and stopping bleeding for the treatment of trauma bleeding, enteritis dysentery, stomatitis, and other diseases [2]. Previous phytochemical studies on P. freyniana afforded 29 compounds, mainly including flavonoids and triterpenoids, but there have been few investigations on biological activities [3].
To further investigate the pharmacologically active constituents of P. freyniana, a phytochemical study on the roots of P. freyniana was conducted, which led to the isolation of two undescribed A-ring contracted triterpenoids (1 and 2) and one new ent-kaurane diterpenoid (22) along with 20 known compounds (Figure 1) from the CH2Cl2-soluble layer. All isolates (1–23) were evaluated for their cytotoxic activities against RA-FLS cells and human cancer cell lines, including HepG2 (human liver carcinoma), HCT-116 (human colorectal carcinoma), BGC-823 (human gastric carcinoma), and MCF-7 (human breast carcinoma), to explore biologically significant metabolites. Herein we deal with the isolation and structure elucidation of the new compounds (1–2, 22) as well as their anti-proliferative activity against RA-FLS cells and cytotoxic activities against the human cancer cell lines Hep-G2, HCT-116, BGC-823, and MCF-7.

2. Results

2.1. Structural Elucidation

Compound 1 was obtained as a white and amorphous powder, and the molecular formula was determined to be C30H46O3 by HR-ESI-MS (m/z 477.3342 [M + Na]+, calculated for C30H46NaO3, 477.3345), indicating eight degrees of unsaturation (Figure S3). The IR spectrum showed characteristic absorptions (Figure S2) of hydroxyl (3325 cm−1), methyl (2928, 2865, 1462, and 1381 cm−1), double bond (1605 cm−1), and ester carbonyl groups (1694 cm−1). In the 1H-NMR spectrum, seven tertiary methyls at δH 0.87, 0.91, 0.95, 0.96, 1.04, 1.16, and 1.20 (each 3H, s); an oxygen-bearing methylene group at δH 4.16 (dd, J = 14.5, 1.8 Hz, 1H) and 4.09 (dd, J = 14.5, 1.7 Hz, 1H); and two olefinic protons at δH 5.40 (s, 1H) and 5.25 (t, J = 3.5 Hz, 1H) were observed (Table 1). The 13C-NMR data with the aid of the DEPT-135° spectrum revealed 30 carbons attributed to seven methyls, nine methylenes, five methines (two olefinic carbons at δC 135.4 and 123.6), and nine quaternary carbons (two olefinic carbons at δC 156.0 and 145.7, and one carbonyl carbon at δC 181.9). Additionally, the degree of unsaturation of 1 and the presence of eight terminal carbons indicated a pentacyclic system containing two double bonds and one carboxyl group with a contracted A-ring rather than a seco structure. The 1D-NMR signals of 1 (Table 1) were similar to those of A-ring contracted oleanane sculponeatic acid (5), except that there was no resonance corresponding to a hydroxyl group on C-19 (δC: 47.0) in 1 [4]. The 1H-1H COSY spectrum showed the presence of four independent spin systems (H-5/H2–6/H2-7, H-9/H2-11/H-12, H2-15/H2-16, and H2-21/H2-22), and interpretation of the HMBC of H2-1 with C-2 and C-3; H-12 with C-11 and C-18; H3-23 with C-24, C-3, C-4, and C-5; H3-24 with C-23, C-4, and C-5; H3-25 with C-2, C-5, C-9, and C-10; H3-26 with C-7 and C-8; H3-27 with C-13, C-14, and C-15; and H3-29 with C-19, C-20, C-21, and C-30 finally confirmed the planar structure of 1 (Figure 2a). To the best of our knowledge, this type of triterpenoid was rarely reported before.
The relative configuration of 1 was ascertained through interpretation of its ROESY spectrum (Figure 2b). Based on the NOE correlations of H-5 with H3-9 and H3-23 and H-9 with H3-27, these protons were classified as α-oriented; the correlations of H3-25 with H3-24 and H3-26 and H-18 with H3-30 revealed their β-orientation. To further elucidate its absolute configuration, the electronic circular dichroism (ECD) spectrum of 1 was recorded in MeOH, and it showed a good agreement with the calculated ECD data of the (5S, 8R, 9S, 10S, 14S, 17S, 18R) model (Figure 3), which supported that the absolute configuration of 1 should be identical with 5S, 8R, 9S, 10S, 14S, 17S, 18R. Hence, the structure of 1 was established as 2-hydroxymethyl-1-norolean-2,12-dien-29-oic acid and it was named madengaisu A.
Compound 2 appeared as a white and amorphous powder. The molecular formula C30H46O3 and eight degrees of unsaturation are the same as those of 1, based on HR-ESI-MS (m/z 477.3342 [M + Na]+, calculated for C30H46NaO3, 477.3361) and 13C NMR data (Figure S12). The NMR data (Table 1) of 2 were almost identical with those of compound 1, except that there was a ursane-type skeleton instead of oleanane; hence, the presence of five tertiary methyls at δH 0.90, 0.95, 1.03, 1.15, and 1.16 (each 3H, s) and two methyl doublets at δH 0.96 (d, J = 4.8 Hz, 3H) and 0.90 (d, J = 6.6 Hz, 3H) in the 1H-NMR spectrum was inferred. Additionally, a H-18/H-19/H-20/H2-21/H2-22 long fragment of the 1H-1H COSY spectrum and HMBC correlations also supported the planar structure of 2 (Figure 4). Its stereochemistry is similar to that of 1, and the β-orientation of H3-29 was deduced from the ROESY cross-peaks between H-18 and H3-29 (Figure 4). In the experimental ECD data, 2 showed a positive cotton effect at 203.4 nm and a negative cotton effect at 222.6 nm, which was consistent with the calculated ECD data of the (4S, 5S, 8R, 9S, 10R, 14S, 17S, 18R, 19S, 20R) model (Figure 5). Thus, the structure of 2 was established as 2-hydroxymethyl-1-norursa-2,12-dien-29-oic acid and it was named madengaisu B accordingly.
Compound 22 was purified as a white amorphous powder. It gave the molecular formula C28H38O6, with 10 degrees of unsaturation, based on HR-ESI-MS (m/z 493.2556 [M + Na]+, calculated for C28H38NaO6, 493.2566). The 13C-NMR and DEPT 135° spectra of 22 displayed 28 carbon resonances, including three methyls (one methoxyl at δC 56.4), ten methylenes, seven methines (three olefinic carbons at δC 113.5, 116.0, and 125.0), and eight quaternary carbons (three olefinic carbons at δC 122.8, 148.8, and 152.9 and two carbonyl carbons at δC 168.4 and 182.4). In addition, two methyl singlets at δH 0.99 and 1.19 (each 3H, s) were apparent in the 1H-NMR spectrum. The above 1D-NMR signals (Table 2) were similar to those of an ent-kaurane type diterpenoid, (-)-17-hydroxy-16α-ent-kauran-19-oic acid [5]. However, one methoxyl singlet at δH 3.90 (3H, s) and three aromatic protons at δH 6.85 (1H, d, J = 8.8 Hz), 7.54 (1H, d, J = 2.0 Hz), and 7.55 (1H, dd, J = 8.7, 2.0 Hz) attributed to a vanilloyl unit were also observed, suggesting that 22 contains a 1,3,4-trisubstituted phenyl ring. Furthermore, the spin system of H2-15/H-16/H-17 deduced from the 1H-1H COSY cross-peaks along with the HMBC correlations from H-2′, H2-17 to C-7′, and H2-17 to C-13 and C-15 illustrated that the hydroxyl at C-17 of the ent-kaurane skeleton was esterified by a vanilloyl group of 22. Thus, the planar structure of 22 is shown in Figure 6. The ECD spectrum (Figure 7) of 22 was also recorded using the TDDFT method to give a result coincident with the experimental spectrum, possessing the absolute configuration of 4R, 5S, 8S, 9R, 10S, 13R, 16R. Therefore, the structure of 22 was identified as 17-(4′-hydroxy-3′-methoxybenzoate)-kaur-16-en-19-oic acid, and it was given the name madengaisu C.
The structures of the following twenty known compounds were identified by comparison of spectroscopic data with the reported literature and ESI-MS analyses: rosamultic acid (3) [6], hyptadienic acid (4) [4], sculponeatic acid (5) [4], ursolic acid (6) [7], pomolic acid (7) [8], euscaphic acid (8) [9], tormentic acid (9) [10], 2-oxo-pomolic acid (10) [11], 3-Hydroxy-13,28-epoxyurs-11-en-28-one (11) [12], cecropiacic acid (12) [13], potentillanoside E (13) [14], rosamultin (14) [15], rubuside A (15) [16], 2α,3β-dihydroxyolean-13(18)-en-28-oic acid (16) [17], camaldulenic acid (17) [18], taraxerol (18) [19], 2α,3β-dihydroxyolean-12-en-28-oic acid 28-O-β-d-glucopyranoside (19) [20], arjunetin (20) [21], alphitolic acid (21) [22], and (-)-kaur-16-en-19-oic acid (23) [23].
Previous reports have shown that an A-ring contracted skeleton was found in ursane-, oleanane-, and dammarane-type triterpenes. In general, the O-containing groups at the C-2 or C-3 position of the six-membered A-ring go through oxidation reactions, which leads to selective C(2)-C(3) bond cleavage and further recyclization to form an A-pentacycle triterpene [24]. In our study, compounds 24 belong to A-ring contracted type triterpenes, which may derive from ursolic acid (6) and corresponding derivatives, and their intermediate products include compounds 712. The plausible biosynthetic pathway of compounds 24 and 612 is shown in Figure 8, and the oleanane-type triterpenes (1, 5, and 16) also have a similar pathway. Meanwhile, aglycones are easily glycosylated at C-28 to form relevant triterpenoid saponins (13, 14, 19, and 20) [25].

2.2. Biological Activities

The cytotoxic activities of all compounds (123) against RA-FLS cells and human cancer cell lines (Hep-G2, HCT-116, BGC-823, and MCF-7) were evaluated in vitro using the MTT method. As shown in Table 3, 6 and 7 showed good bioactivities against RA-FLS cells with IC50 values of 24.63 ± 1.96 and 25.12 ± 1.97 μM, respectively. Among the tested triterpenes, 4 and 19 exhibited a moderate cytotoxic effect on Hep-G2 cells (25.16 ± 2.55 and 17.66 ± 1.82 μM, respectively), and 16 and 21 were observed to inhibit HCT-116 cells (13.25 ± 1.65 and 21.62 ± 0.33 μM, respectively). Furthermore, 2 and 16 showed cytotoxic activities against BGC-823 with IC50 values of 24.76 ± 0.94 and 26.83 ± 2.52 μM, respectively. Except for the compounds mentioned above, other compounds with IC50 values higher than 30 μM were considered to be inactive against RA-FLS cells and human cancer cell lines.

3. Materials and Methods

3.1. General Information

Recording of 1D-NMR and 2D-NMR spectra was performed on a Bruker AV-600 spectrometer (Bruker, Billerica, MA, USA) with a single NMR probe at 600 MHz for 1H and 150 MHz for 13C in CD3OD. HR-ESI-MS experiments were performed using an Agilent 6200 series TOF/6500 series (Agilent, Santa Clara, CA, USA). A PerkinElmer Frontier MIR spectrometer (PerkinElmer, Waltham, MA, USA) was used to determine IR spectra. UV spectra were recorded on a PerkinElmer Lambda 650 (PerkinElmer, Waltham, MA, USA) in methanol. Optical rotations of compounds were determined by a Rudolph Research Analytical Autopol IV automatic polarimeter (Rudolph, Hackettstown, NJ, USA). An Applied Photophysics Chirascan plus CD spectrometer was used to determine ECD spectra. Semi-preparative RP-HPLC (Agilent, Palo Alto, CA, USA) was carried out on an Agilent 1100 system with an Agilent Eclipse XDB-C18 column (5 μm, 4.6 × 250 mm, Agilent, Palo Alto, CA, USA). Sephadex LH-20 for column chromatography was obtained from Pharmacia Fine Chemical Company, Ltd. (Uppsala, Sweden). Column chromatographic silica gel (80-100 and 200-300 mesh) and TLC plates (GF254) were purchased from Qingdao Marine Chemical Inc. (Qingdao, China). Acetonitrile and methanol (HPLC-grade) were obtained from Merck KGaA (Darmstadt, Germany). The other solvents were purchased from Shanghai Titan Scientific Co., Ltd. (Shanghai, China).

3.2. Plant Material

The dried roots of P. freyniana were collected at Huaihua of Hunan Province, China, in October 2019 and authenticated by Zai-Qi Zhang, a senior professor at Hunan University of Medicine. A voucher specimen (201910PF) was deposited at the TCM and Ethnomedicine Innovation and Development International Laboratory, School of Pharmacy, Hunan University of Chinese Medicine, Changsha, Hunan, People’s Republic of China.

3.3. Extraction and Isolation

The air-dried and pulverized roots of P. freyniana (10 kg) were extracted with 95% EtOH (10 L × 3, each 7 d). After concentration under reduced pressure, 600 g of crude EtOH extract was obtained. The EtOH extract was suspended in H2O and successively partitioned with PE (petroleum ether), CH2Cl2, EtOAc, and n-BuOH. Among them, the CH2Cl2-soluble fraction (41.5 g) was subjected to silica gel column chromatography (CC) eluted with PE/EtOAc (1:0–0:1) gradients to afford fourteen major fractions (C1–C14). Fraction C5 (587.4 mg) was fractioned on a Sephadex LH-20 column (CH3Cl/CH3OH, 1:1, v/v) to give three subfractions (C5.1–C5.3); then, subfraction C5.1 (387.4 mg) was isolated and purified by silica gel CC (gradient system, PE/EtOAc, 1:0–20:1, v/v) to obtain 18 (2.3 mg). Fraction C7 (1.2 g) was separated by a Sephadex LH-20 column with eluent (CH3Cl/CH3OH, 1:1, v/v) to give four subfractions (C7.1–C7.4). Subfraction C7.4 (38.8 mg) was chromatographed on a silica gel column (PE/EtOAc, 50:1–20:1, v/v) to give 23 (13.0 mg). Fraction C10 (1.2 g) was subjected to silica gel CC (PE/EtOAc, 20:1–8:1, v/v) after being separated on a Sephadex LH-20 column eluted with CH3Cl/CH3OH (1:1, v/v) to give nine subfractions (C10.1–C10.9). Subfraction C10.6 (99.6 mg) was purified by semi-preparative HPLC (CH3CN/CH3OH/H2O, 90:8:2) to obtain 1 (2.3 mg, tR 9.8 min), 2 (8.3 mg, tR 10.5 min), and 6 (36.0 mg, tR 13.5 min). Subfraction C10.8 (140.5 mg) was purified by semi-preparative HPLC with the mobile phase (CH3CN/H2O, 87:13) after being separated with silica gel CC eluted with PE/EtOAc (20:1–8:1, v/v) gradients to yield 11 (1.0 mg, tR 16.8 min). Fraction C11 (1.8 g) was chromatographed with a silica gel column (PE/EtOAc, 1:0–0:1, v/v) to give twelve subfractions (C11.1–C11.12), and subfraction C11.8.3 (233.6 mg) was purified by semi-preparative HPLC using CH3OH/CH3CN/H2O (40:40:20–45:45:10) as a gradient eluent to provide 5 (21.5 mg, tR 14.7 min). Similarly, C11.10 (135.5 mg) was isolated by semi-preparative HPLC (CH3OH/H2O, 80:20–100:0) to give four further subfractions (C11.10.1–C11.10.4), and 7 (42.8 mg, tR 11.6 min) was yielded by semi-preparative HPLC, eluting with the gradient mobile phase CH3OH/CH3CN/H2O (41:41:18–42:42:16) from subfraction C11.10.3 (38.1 mg). Then, the purification of 22 (2.9 mg, tR 7.9 min) was conducted by semi-preparative HPLC (CH3CN/H2O, 90:10) from C11.10.4 (4.9 mg). Furthermore, subfraction C11.11 (485.7 mg) was isolated with silica gel CC (PE/EtOAc, 50:3, v/v) before being purified by semi-preparative HPLC using CH3CN/H2O (70:30–100:0) as the mobile phase to obtain 10 (0.9 mg, tR 7.4 min). Fraction C12 (7.9 g) was subjected to silica gel CC using 2% dichloromethane in methanol as an isocratic eluent to afford five subfractions (C12.1–C12.5), and compound 12 (6.2 mg, tR 6.4 min) was purified by semi-preparative HPLC (CH3OH/H2O, 82:18) from subfraction C12.4 (179.1 mg). C12.2 (3.0 g) was fractioned with silica gel CC (CH2Cl2/CH3OH, 49:1, v/v) to afford five subfractions (C12.2.1–C12.2.5). Among them, subfraction C12.2.3 (618.6 mg) was separated using silica gel CC (CH2Cl2/CH3OH, 99:1, v/v) and then by semi-preparative HPLC using 83% CH3OH in H2O as an isocratic eluent to yield 4 (34.1 mg, tR 12.6 min). Subfraction C12.2.4 (408.3 mg) was further subjected to a Sephadex LH-20 column with eluent (CH3Cl/CH3OH,1:1, v/v) before being isolated by semi-preparative HPLC (CH3OH/H2O, 80:20–90:10) to give 8 (8.8 mg, tR 8.6 min), 17 (3.0 mg, tR 15.2 min), and two more subfractions (C12.2.4.1–C12.2.4.2). In addition, compounds 21 (5.0 mg, tR 6.7 min) and 16 (4.5 mg, tR 7.4 min) were obtained by semi-preparative HPLC (CH3CN/H2O, 98:2) from subfraction C12.2.4.2 (15.1 mg). Fraction C14 (11.2 g) was chromatographed with a silica gel column (CH2Cl2/CH3OH, 10:1, v/v) to afford three subfractions (C14.1–C14.3). Subfraction C14-1 (5.7 g) was further subjected to silica gel CC to obtain five subfractions (C11.14.1–C11.14.5), and C14.1.4 (871.5 mg) was purified by silica gel CC eluted with CH2Cl2/CH3OH (10:1, v/v) to yield 3 (261.4 mg). Then, C14.1.5 (207.3 mg) was isolated by semi-preparative HPLC (CH3CN/H2O 35:65–60:40) after being separated by silica gel CC (CH2Cl2/CH3OH, 47:3, v/v) to provide 20 (2.6 mg, tR 7.3 min) and two further subfractions (C14.1.5.1-C14.1.5.2). C14.1.5.2 (5.0 mg) was isolated by semi-preparative HPLC using 28% CH3OH in H2O to yield 19 (2.5 mg, tR 21.2 min) and 15 (1.8 mg, tR 23.9 min). Similarly, compounds 9 (157.3 mg) and 14 (49.3 mg) were obtained by silica gel CC (CH2Cl2/CH3OH, 20:1, v/v) from subfraction C14.3 (1.5 g).

3.4. Characterization

2-hydroxymethyl-1-norolean-2,12-dien-29-oic acid (1): White amorphous powder; [ α ] D 25 = +40.0 (c 0.1, MeOH); ECD (MeOH) Δε (nm): +26.2 (210) and -28.9 (220); IR (KBr) υmax: 3325, 2928, 2865, 1694, 1605, 1462, 1381, and 667 cm−1; for 1H- and 13C-NMR data, see Table 1; HR-ESI-MS (Pos.) m/z 477.3342 [M + Na]+ (calcd for C30H46NaO3, 477.3345).
2-hydroxymethyl-1-norursa-2,12-dien-29-oic acid (2): White amorphous powder; [ α ] D 25 = +50.0 (c 0.1, MeOH); ECD (MeOH) Δε (nm): +66.3 (203) and −11.7 (223); IR (KBr) υmax: 3362, 2964, 2860, 1687, 1585, 1459, 1380, and 1023 cm−1; for 1H- and 13C-NMR data, see Table 1; HR-ESI-MS (Pos.) m/z 477.3361 [M + Na]+ (calcd for C30H46NaO3, 477.3345).
17-(4′-hydroxy-3′-methoxybenzoate)-kaur-16-en-19-oic acid (22): White amorphous powder; [ α ] D 25 = −80.0 (c 0.1, MeOH); UV (MeOH) λmax(log ε) = 199 (4.5) nm; ECD (MeOH) Δε (nm): +19.6 (209) and −27.2 (223); IR (KBr) υmax: 3406, 2921, 1698, 1280, and 1030 cm−1; for 1H- and 13C-NMR data, see Table 2; HR-ESI-MS (Pos.) m/z 493.2556 [M + Na]+ (calcd for C28H38NaO6493.2566).

3.5. Cell Culture

The human RA-FLS cell line and human cancer cell lines (Hep-G2, HCT-116, BGC-823, and MCF-7) were purchased from Shanghai Fu-Heng Biological Technology Co., Ltd. (Shanghai, China). RA-FLS cells were cultured in DMEM/F-12 with 10% fetal bovine serum (FBS); the Hep-G2, BGC-823, and MCF-7 cell lines in DMEM with 10% FBS; and the HCT-116 cell line in McCoy’s 5A with 10% FBS. Cell cultures were maintained at 37 °C in a humidified, 5% CO2 atmosphere [26].

3.6. Cytotoxicity Assessment

The cytotoxicity activities were evaluated by the MTT assay [26]. Firstly, the tested compounds (123) were diluted in corresponding media from freshly made solutions in DMSO (50 mM) to working concentrations (0–40 μM). The cells were seeded in 96-well plates. After 24 h incubation, the cells were exposed to the test compounds at the indicated concentrations for 24 h under 5% CO2 at 37 °C. Then, 100 μL of MTT (0.5 mg/mL) was added to each well and incubated for 4 h. Subsequently, the supernatant was removed from the formazan crystals, and 100 μL DMSO was added to each well. Finally, the absorbance was measured by a microplate reader using a wavelength of 492 nm to determine cell viability rate.

4. Conclusions

As a common Dong ethnomedicine in China, the roots of P. freyniana have been widely used for heat clearing and detoxifying. Moreover, ursane-type triterpenes are some of the most common pentacyclic triterpenes in ethnic medicine, with various pharmacological activities and low toxicity [27]. In this work, twenty-three compounds, comprising two unusual A-ring contracted triterpenoids named madengaisu A and B (1 and 2) and one new ent-kaurane diterpenoid named madengaisu C (22) along with 20 known compounds, were isolated and identified, thirteen of which were classified as ursane-type triterpenoids. In addition, this article is the first to report the diterpenoid constituents from the genus Potentilla L. Compounds 5, 11, 15, 16, 19, 22, and 23 were isolated from the genus Potentilla L. for the first time, and compounds 4, 6, 7, 10, and 20 were isolated from this plant for the first time, which enriched the chemical metabolite diversity of diterpenoids and triterpenoids in this plant. Additionally, compounds 2, 4, 6, 7, 16, 19, and 21 showed good anti-RA and/or anti-tumor activities, which demonstrated that triterpenes from P. freyniana may serve as therapeutic agents for RA and cancer treatment.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27123665/s1, Figures S1–S26: UV, IR, HR ESIMS and NMR spectra of compound 12 and 22.; Tables S1–S6: ECD calculations results of compound 12 and 22. Refs. [28,29,30] are cited in Supplementary Materials.

Author Contributions

Conceptualization, W.W.; data curation, X.-D.Z., L.-L.C. and B.L.; formal analysis, Z.-L.C., H.-H.Y. and Y.-P.Y.; funding acquisition, Z.-Q.Z., B.L. and W.W.; investigation, J.W. and B.L.; methodology, H.-H.Y. and B.L.; project administration, W.W.; resources, W.W.; supervision, Z.-Q.Z.; validation, Q.-Y.Y.; visualization, J.W. and Z.-L.C.; writing—original draft, J.W.; writing—review and editing, L.-L.C. and B.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the Hunan Provincial Key Laboratory Project of Dong Medicine Research (grant number 2015TP1020-04), the Hunan Provincial Key Laboratory Platform Construction Project of Dong Medicine Research (grant number 2017CT5025), the Open Project of the Hunan Provincial Key Laboratory of Dong Medicine Research (grant number YS2019YFE012440), Natural Science Foundation of Hunan province (grant number 2020JJ4463); Research Foundation of Education Bureau of Hunan Province (grant number 20A367), and the Scientific project of the Hunan Provincial Health Commission (grant number 202113052008).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The raw data that support the findings of this study are available from the corresponding authors (J.W., L.B., and W.W.) upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds 123 are available from the authors.

References

  1. Augustynowicz, D.; Latté, K.P.; Tomczyk, M. Recent phytochemical and pharmacological advances in the genus Potentilla L. sensu lato—An update covering the period from 2009 to 2020. J. Ethnopharmacol. 2021, 266, 113412. [Google Scholar] [CrossRef]
  2. Wu, J.; Zhang, Z.-Q.; Yu, H.-H.; Huang, F.-B.; Chen, Z.-L.; Chu, L.-L.; Li, B.; Wang, W. Research progress on chemical constituents and pharmacological activities of Potentilla. China J. Chin. Mater. Med. 2022, 47, 1509–1538. [Google Scholar]
  3. Huang, F.-B.; Wu, J.; Sheng, W.-B.; Peng, C.-Y.; Wang, W.; Zhang, Z.-Q.; Li, B.; Chen, S.-H. Research Progress on Chemical Constituents and Pharmacological Activities of Potentilla freyniana Bornm. J. Hunan Univ. Chin. Med. 2020, 40, 1039–1044. [Google Scholar]
  4. Wang, F.; Li, X.-M.; Liu, J.-K. New terpenoids from Isodon sculponeata. Chem. Pharm. Bull. 2009, 57, 525–527. [Google Scholar] [CrossRef] [Green Version]
  5. Wu, Y.-C.; Hung, Y.-C.; Chang, F.-R.; Cosentino, M.; Wang, H.-K.; Lee, K.-H. Identification of ent-16β, 17-dihydroxykauran-19-oic acid as an anti-HIV principle and isolation of the new diterpenoids annosquamosins A and B from Annona squamosa. J. Nat. Prod. 1996, 59, 635–637. [Google Scholar] [CrossRef]
  6. Yeo, H.; Park, S.-Y.; Kim, J. A-ring contracted triterpenoid from Rosa multiflora. Phytochemistry 1998, 48, 1399–1401. [Google Scholar] [CrossRef]
  7. Seebacher, W.; Simic, N.; Weis, R.; Saf, R.; Kunert, O. Complete assignments of 1H and 13C NMR resonances of oleanolic acid, 18α-oleanolic acid, ursolic acid and their 11-oxo derivatives. Magn. Reson. Chem. 2003, 41, 636–638. [Google Scholar] [CrossRef]
  8. Liu, W.-J.; Hou, X.-Q.; Chen, H.; Liang, J.-Y.; Sun, J.-B. Chemical constituents from Agrimonia pilosa Ledeb. and their chemotaxonomic significance. Nat. Prod. Res. 2016, 30, 2495–2499. [Google Scholar] [CrossRef]
  9. Zheng, G.; Piao, H. Study on chemical constituents from Potentilla supina. Zhong Cao Yao 2012, 43, 1285–1288. [Google Scholar]
  10. Reher, G.; Buděšínský, M. Triterpenoids from plants of the Sanguisorbeae. Phytochemistry 1992, 31, 3909–3914. [Google Scholar] [CrossRef]
  11. Song, N.-Y.; Cho, J.-G.; Im, D.; Lee, D.-Y.; Wu, Q.; Seo, W.-D.; Kang, H.C.; Lee, Y.-H.; Baek, N.-I. Triterpenoids from Fragaria ananassa calyx and their inhibitory effects on melanogenesis in B16-F10 mouse melanoma cells. Nat. Prod. Res. 2013, 27, 2219–2223. [Google Scholar] [CrossRef] [PubMed]
  12. Al Musayeib, N.M.; Mothana, R.A.; Gamal, A.A.E.; Al-Massarani, S.M.; Maes, L. In vitro antiprotozoal activity of triterpenoid constituents of Kleinia odora growing in Saudi Arabia. Molecules 2013, 18, 9207–9218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Huang, X.-Y.; Ma, G.-X.; Zhong, X.-Q.; Zhou, Y.-L.; Dai, H.-N.; Wu, H.-F.; Zhu, Y.-D.; Yang, J.-S.; Yuan, J.-Q.; Xu, X.-D. Triterpene constituents from Rosa cymosa Tratt. China J. Chin. Mater. Med. 2014, 39, 4637–4641. [Google Scholar]
  14. Morikawa, T.; Ninomiya, K.; Imura, K.; Yamaguchi, T.; Akagi, Y.; Yoshikawa, M.; Hayakawa, T.; Muraoka, O. Hepatoprotective triterpenes from traditional Tibetan medicine Potentilla anserina. Phytochemistry 2014, 102, 169–181. [Google Scholar] [CrossRef]
  15. LI, Y.-L. A new triterpenic acid from roots of Rosa multiflora var. cathayensis. Chin. Tradit. Herb. Drugs. 2017, 48, 4208–4214. [Google Scholar]
  16. Li, W.; Fu, H.; Bai, H.; Sasaki, T.; Kato, H.; Koike, K. Triterpenoid saponins from Rubus ellipticus var. obcordatus. J. Nat. Prod. 2009, 72, 1755–1760. [Google Scholar] [CrossRef] [PubMed]
  17. Zeng, N.; Shen, Y.; Li, L.-Z.; Jiao, W.-H.; Gao, P.-Y.; Song, S.-J.; Chen, W.-S.; Lin, H.-W. Anti-inflammatory triterpenes from the leaves of Rosa laevigata. J. Nat. Prod. 2011, 74, 732–738. [Google Scholar] [CrossRef] [PubMed]
  18. Begum, S.; Farhat, a.; Siddiqui, B.S. Triterpenoids from the leaves of Eucalyptus camaldulensis var. obtusa. J. Nat. Prod. 1997, 60, 20–23. [Google Scholar] [CrossRef]
  19. Wiedemann, B.; Lerche, H.; Lotter, H.; Neszmelyi, A.; Wagner, H.; Müller, A.A. Two novel triterpenoids from the stemwood of Herrania cuatrecasana. Phytochemistry 1999, 52, 333–337. [Google Scholar] [CrossRef]
  20. Tabopda, T.K.; Ngoupayo, J.; Tanoli, S.A.K.; Mitaine-Offer, A.-C.; Ngadjui, B.T.; Ali, M.S.; Luu, B.; Lacaille-Dubois, M.-A. Antimicrobial pentacyclic triterpenoids from Terminalia superba. Planta Med. 2009, 75, 522–527. [Google Scholar] [CrossRef]
  21. Dai, H.-N.; Ma, G.-X.; Zhou, J.-M.; Zhong, X.-Q.; Zhou, Y.-L.; Lv, G.-R.; Wang, Y.-Q.; Yuan, J.-Q.; Xu, X.-D. Triterpenoids from roots of Rosa laevigata. China J. Chin. Mater. Med. 2016, 41, 2267–2272. [Google Scholar]
  22. Ren, Y.; Anaya-Eugenio, G.D.; Czarnecki, A.A.; Ninh, T.N.; Yuan, C.; Chai, H.-B.; Soejarto, D.D.; Burdette, J.E.; de Blanco, E.J.C.; Kinghorn, A.D. Cytotoxic and NF-κB and mitochondrial transmembrane potential inhibitory pentacyclic triterpenoids from Syzygium corticosum and their semi-synthetic derivatives. Bioorg. Med. Chem. 2018, 26, 4452–4460. [Google Scholar] [CrossRef] [PubMed]
  23. Zhang, G.; Shimokawa, S.; Mochizuki, M.; Kumamoto, T.; Nakanishi, W.; Watanabe, T.; Ishikawa, T.; Matsumoto, K.; Tashima, K.; Horie, S. Chemical constituents of Aristolochia constricta: Antispasmodic effects of its constituents in guinea-pig ileum and isolation of a diterpeno−lignan hybrid. J. Nat. Prod. 2008, 71, 1167–1172. [Google Scholar] [CrossRef] [PubMed]
  24. Grishko, V.; Tolmacheva, I.; Pereslavtseva, A. Triterpenoids with a five-membered A-ring: Distribution in nature, transformations, synthesis, and biological activity. Chem. Nat. Compd. 2015, 51, 1–21. [Google Scholar] [CrossRef]
  25. Rao, K.R.; Rao, L.; Rao, N.P. An A-ring contracted triterpenoid from Hyptis suaveolens. Phytochemistry 1990, 29, 1326–1329. [Google Scholar]
  26. Shen, B.; Chen, S.; Zhou, Q.; Jian, Y.; Daniyal, M.; Sheng, W.; Gong, L.; Luo, D.; Liu, B.; Xu, G. Flavonoid glycosides from the rhizomes of Pronephrium penangianum. Phytochemistry 2020, 179, 112500. [Google Scholar] [CrossRef]
  27. Sun, Q.; He, M.; Zhang, M.; Zeng, S.; Chen, L.; Zhou, L.; Xu, H. Ursolic acid: A systematic review of its pharmacology, toxicity and rethink on its pharmacokinetics based on PK-PD model. Fitoterapia 2020, 147, 104735. [Google Scholar] [CrossRef]
  28. Sybyl Software, Version X 2.0; Tripos Associates Inc.: St. Louis, MO, USA, 2013.
  29. Neese, F. The ORCA program system, Wiley Interdiscip. Rev. Comput. Mol. Sci. 2012, 2, 73–78. [Google Scholar] [CrossRef]
  30. Stephens, P.J.; Harada, N. ECD cotton effect approximated by the Gaussian curve and other methods. Chirality 2010, 22, 229–233. [Google Scholar] [CrossRef]
Figure 1. Chemical structures of compounds 1–23.
Figure 1. Chemical structures of compounds 1–23.
Molecules 27 03665 g001
Figure 2. (a) Key 1H-1H COSY and HMBC correlations of 1. (b) Selected ROESY correlations of 1.
Figure 2. (a) Key 1H-1H COSY and HMBC correlations of 1. (b) Selected ROESY correlations of 1.
Molecules 27 03665 g002
Figure 3. Experimental and calculated ECD spectra of 1.
Figure 3. Experimental and calculated ECD spectra of 1.
Molecules 27 03665 g003
Figure 4. (a) Key 1H-1H COSY and HMBC correlations of 2. (b) Selected ROESY correlations of 2.
Figure 4. (a) Key 1H-1H COSY and HMBC correlations of 2. (b) Selected ROESY correlations of 2.
Molecules 27 03665 g004
Figure 5. Experimental and calculated ECD spectra of 2.
Figure 5. Experimental and calculated ECD spectra of 2.
Molecules 27 03665 g005
Figure 6. Key 2D-NMR correlations of compound 22.
Figure 6. Key 2D-NMR correlations of compound 22.
Molecules 27 03665 g006
Figure 7. Experimental and calculated ECD spectra of compound 22.
Figure 7. Experimental and calculated ECD spectra of compound 22.
Molecules 27 03665 g007
Figure 8. Plausible biosynthetic pathway for 24 and 612.
Figure 8. Plausible biosynthetic pathway for 24 and 612.
Molecules 27 03665 g008
Table 1. The 1H and 13C-NMR data of 1 and 2 (600/150 MHz, CD3OD, δ in ppm, J in Hz).
Table 1. The 1H and 13C-NMR data of 1 and 2 (600/150 MHz, CD3OD, δ in ppm, J in Hz).
Position12
δCδHδCδH
161.44.16, dd (14.5, 1.8)
4.09, dd (14.5, 1.7)
61.44.16, dd (14.5, 1.8)
4.08, dd (14.5, 1.7)
2156.0-156.1-
3135.45.40, s135.55.39, s
443.3-43.3-
564.51.44, m64.41.41, m
618.21.52, m18.21.49, m
734.9a 1.56, m
b 1.38, m
35.4a 1.58, m
b 1.39, m
842.1-42.3-
944.92.15, m44.82.09, m
1051.8-51.7-
1127.5b 2.19, m
a 2.03, m
27.4b 2.18, m
a 2.07, m
12123.65.25, t (3.5)126.85.22, t (3.5)
13145.7-140.1-
1443.2-43.5-
1529.1b 1.83, dt (13.8, 4.3)
a 1.09, dt (13.7, 3.4)
29.5b 1.95, td (13.6, 4.5)
a 1.08, ddd (13.7, 4.3, 2.4)
1624.0a 2.05, m
b 1.61, ddt (13.6, 4.2, 2.2)
25.3a 2.05, m b
b 1.65, m a
1747.7-(48.6) *-
1843.02.85, dd (13.9, 4.6)54.52.20, m
1947.0a 1.72, m40.21.38.m
2031.6-40.40.97, overlapped
2135.1a 1.42, m
b 1.23, m
31.71.51, m
1.36, m
2233.8b 1.77, m
a 1.55, m
38.11.68, m
1.63, m
2330.31.04, s30.31.03, s
2421.80.96, s21.80.95, s
2519.31.16, s19.31.16, s
2619.20.87, s19.30.90, s
2726.81.20, s24.31.15, s
28181.9-181.7-
2933.60.91, s17.70.90, d (6.6)
3024.00.95, s21.60.96, d (4.8)
a: α-oriented proton. b: β-oriented proton. * Overlapped with the solvent peak.
Table 2. The 1H and 13C-NMR data of compound 22 (600/150 MHz, CD3OD, δ in ppm, J in Hz).
Table 2. The 1H and 13C-NMR data of compound 22 (600/150 MHz, CD3OD, δ in ppm, J in Hz).
Position22Position22
δCδHδCδH
142.21.91, m
0.86, m
1546.21.67, overlapped
1.08, m
220.41.92, m
1.30, m
1641.22.26, m
339.32.13, overlapped
1.01, m
1769.94.09, d (1.7)
4.08, d (1.7)
444.8-1829.61.19, s
558.31.07, overlapped19182.4-
623.71.87, m2016.30.99, s
742.91.50, m1′122.8-
846.2-2′113.57.54, d (2.0)
956.81.07, overlapped3′148.8-
1040.8-4′152.9-
1119.81.69, overlapped5′116.06.85, d (8.8)
1232.51.60, m
1.48, m
6′125.07.55, dd (2.0 8.7)
1340.22.16, m7′168.4-
1438.31.95, m
1.16, m
-OCH356.43.90, s
Table 3. Cytotoxic effects of tested compounds on RA-FLS and human cancer cell lines.
Table 3. Cytotoxic effects of tested compounds on RA-FLS and human cancer cell lines.
CompoundIC50 (μM) a
Anti-RA-FLS ActivityAnti-Tumor Activity
RA-FLSHep-G2HCT-116BGC-823MCF-7
2>30 b>30>3024.76 ± 0.94>30
4>3025.16 ± 2.55>30>30>30
624.63 ± 1.96>30>30>30>30
725.12 ± 1.97>30>30>30>30
16>30>3013.25 ± 1.6526.83 ± 2.52>30
19>3017.66 ± 1.82>30>30>30
21>30>3021.62 ± 0.33>30>30
Methotrexate c5.09 ± 0.60----
Taxol (nM) c-13.82 ± 0.7815.79 ± 0.919.12 ± 1.23-
a IC50, the half maximal inhibitory concentration. Values represent the mean ± SD of three independent experiments. b Compounds exhibiting IC50 values > 30 μM were considered to be inactive. c Methotrexate as positive control for anti-RA-FLS activity test and taxol for anti-tumor activity test.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wu, J.; Zhang, Z.-Q.; Zhou, X.-D.; Yao, Q.-Y.; Chen, Z.-L.; Chu, L.-L.; Yu, H.-H.; Yang, Y.-P.; Li, B.; Wang, W. New Terpenoids from Potentilla freyniana Bornm. and Their Cytotoxic Activities. Molecules 2022, 27, 3665. https://doi.org/10.3390/molecules27123665

AMA Style

Wu J, Zhang Z-Q, Zhou X-D, Yao Q-Y, Chen Z-L, Chu L-L, Yu H-H, Yang Y-P, Li B, Wang W. New Terpenoids from Potentilla freyniana Bornm. and Their Cytotoxic Activities. Molecules. 2022; 27(12):3665. https://doi.org/10.3390/molecules27123665

Chicago/Turabian Style

Wu, Jia, Zai-Qi Zhang, Xu-Dong Zhou, Qing-Ying Yao, Zhu-Liang Chen, Ling-Ling Chu, Huang-He Yu, Yu-Pei Yang, Bin Li, and Wei Wang. 2022. "New Terpenoids from Potentilla freyniana Bornm. and Their Cytotoxic Activities" Molecules 27, no. 12: 3665. https://doi.org/10.3390/molecules27123665

Article Metrics

Back to TopTop