Next Article in Journal
Phytochemical Constituent Analysis of Phyllanthus emblica L. Fruit Nanoherbals by LC-HRMS and Their Antimutagenic Activity and Teratogenic Effects
Previous Article in Journal
A Review on the Design of Carbon-Based Nanomaterials as MRI Contrast Agents
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Triterpenoids from the Leaves of Diospyros digyna and Their PTP1B Inhibitory Activity

1
Guangdong Provincial Engineering Research Center for Modernization of TCM, Jinan University, Guangzhou 510632, China
2
NMPA Key Laboratory for Quality Evaluation of TCM, Jinan University, Guangzhou 510632, China
3
Guangdong Institute for Drug Control, Guangzhou 510663, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2024, 29(7), 1640; https://doi.org/10.3390/molecules29071640
Submission received: 14 March 2024 / Revised: 2 April 2024 / Accepted: 3 April 2024 / Published: 5 April 2024
(This article belongs to the Section Natural Products Chemistry)

Abstract

:
Six new 2α-hydroxy ursane triterpenoids, 3α-cis-p-coumaroyloxy-2α,19α-dihydroxy-12-ursen-28-oic acid (1), 3α-trans-p-coumaroyloxy-2α,19α-dihydroxy-12-ursen-28-oic acid (2), 3α-trans-p-coumaroyloxy-2α-hydroxy-12-ursen-28-oic acid (3), 3β-trans-p-coumaroyloxy-2α-hydroxy-12,20(30)-ursadien-28-oic acid (4), 3β-trans-feruloyloxy-2α-hydroxy-12,20(30)-ursadien-28-oic acid (5), and 3α-trans-feruloyloxy-2α-hydroxy-12,20(30)-ursadien-28-oic acid (6), along with eleven known triterpenoids (717), were isolated from the leaves of Diospyros digyna. Their chemical structures were elucidated by comprehensive analysis of UV, IR, HRESIMS, and NMR spectra. All the isolated compounds were evaluated for their PTP1B inhibitory activity. 3β-O-trans-feruloyl-2α-hydroxy-urs-12-en-28-oic acid (13) showed the best inhibition activity with an IC50 value of 10.32 ± 1.21 μM. The molecular docking study found that the binding affinity of compound 13 for PTP1B was comparable to that of oleanolic acid (positive control).

Graphical Abstract

1. Introduction

Type 2 diabetes (T2DM), a chronic metabolic disease, primarily results from an impaired insulin receptor signaling pathway [1]. T2DM accounts for approximately 90% of cases of diabetes. The multifactorial etiology of T2DM, including genetic, environmental, and lifestyle factors, has made its management and prevention a significant challenge for healthcare systems worldwide [2]. Protein tyrosine phosphatase 1B (PTP1B), an intracellular enzyme, has been implicated in the negative regulation of insulin signaling [3]. The overexpression of PTP1B in various tissues of T2DM patients highlights its pivotal role in the pathogenesis of insulin resistance [4]. Recent studies have demonstrated that genetic deletion or the pharmacological inhibition of PTP1B enhances insulin sensitivity and protects against diet-induced obesity [5,6,7,8], thereby underscoring PTP1B as a crucial therapeutical target for T2DM. Several PTP1B inhibitors have shown promise in preclinical models by improving glycemic control and insulin sensitivity [9]. Natural products are a rich source of new drug candidates [10] and an important source of PTP1B inhibitors [11,12].
The Diospyros genus comprises over 500 species of evergreen trees and shrubs, belonging to the family of Ebenaceae [13]. Diospyros plants are widely distributed in pantropical regions, and some species (particularly Diospyros digyna Jacq.) are edible fruit-yielding plants [14]. The chemical constituents of the plants, including triterpenoids, flavonoids, tannins, sugars, and phenolic acids, show antioxidant, anti-inflammatory, antiviral, antitumor, and PTP1B inhibitory activities [15,16,17]. In the search for natural PTP1B inhibitors, a phytochemical investigation on the leaves of D. digyna was carried out. Six new 2α-hydroxy ursane triterpenoids (16), along with eleven known ones were isolated from this plant (Figure 1). In addition, the PTP1B inhibitory activity of these triterpenoids was tested. Herein, the isolation, structural elucidation, and bioactivity of the isolates are presented.

2. Results

Compound 1 was isolated as a white amorphous powder. Its molecular formula was deduced as C39H54O7 based on its HRESIMS ion at m/z 635.3934 [M + H]+. The UV spectrum showed the absorption maxima at 206, 228, and 310 nm. The IR spectrum suggested the presence of hydroxy (3425 cm−1), carbonyl (1694 cm−1), and aromatic (1605, 1513, and 1454 cm−1) groups. The 1H and 13C NMR spectra showed a cis-p-coumaroyl group [δH 7.66 (2H, d, J = 8.7 Hz), 6.88 (1H, d, J = 13.0 Hz), 6.75 (2H, d, J = 8.7 Hz), 5.87 (1H, d, J = 13.0 Hz); δC 168.6, 160.0, 144.8, 133.8 (×2), 127.9, 117.5, 116.0 (×2)], an olefinic bond [δH 5.32 (1H, m); δC 140.2, 129.4], two oxymethines [δH 5.00 (1H, d, J = 4.1 Hz), 4.11 (1H, dt, J = 11.0, 4.1 Hz); δC 81.4, 66.2], and seven methyls [δH 1.36 (3H, s), 1.23 (3H, s), 1.04 (3H, s), 1.00 (3H, s), 0.95 (3H, d, J = 6.7 Hz), 0.91 (3H, s), 0.81 (3H, s); δC 28.7, 27.2, 25.1, 22.4, 17.7, 17.0, 16.8] (Table 1). Comparison of the 1D NMR spectra of 1 with those of 3-O-cis-p-coumaroyltormentic acid [18] indicated similar planar structures, which was further verified by the 2D NMR spectra of 1.
The 1H–1H COSY spectrum of 1 suggested spin-coupling systems of H2-1/H-2/H-3, H-5/H2-6/H2-7, H-9/H2-11/H-12, H2-15/H2-16, H3-30/H-20/H2-21/H2-22, H-2′/H-3′, H-5′/H-6′, and H-8′/H-9′ (Figure 2). In the HMBC spectrum, the correlations from H2-1/H-3/H2-7/H-9/H3-25 to C-5, from H-9/H-12/H-18 to C-14, from H-18 to C-12/C-14/C-16/C-28/C-29, from H-20/H-22 to C-18, from H3-23 to C-3/C-5/C-24, from H3-25 to C-1/C-5/C-9, from H3-26 to C-7/C-9, from H3-27 to C-8/C-13/C-15, and from H3-30 to C-19/C-21, established an ursane-type triterpenoid skeleton (Figure 2). Moreover, the HMBC correlation from H-3 to C-1′ located the cis-p-coumaroyl group at C-3. In the NOESY spectrum, the correlations between H-9/H3-24 and H-5, between H-9/H-16α and H3-27, and between H-16α and H3-30 suggested H-5, H-9, H3-24, and H3-27 were α-oriented. The NOE correlations between H-2/H3-23/H3-26 and H3-25, between H-3 and H3-23, and between H-20/H3-29 and H-18 suggested H-2, H-3, H-18, H-20, H3-23, H3-25, H3-26, and H3-29 were β-oriented (Figure 3). Thus, the structure of 1 was elucidated and named as 3α-cis-p-coumaroyloxy-2α,19α-dihydroxy-12-ursen-28-oic acid.
The molecular formula of 2 was determined as C39H54O7 by the HRESIMS at m/z 635.3938 [M + H]+ (calcd for C39H55O7, 635.3942). The UV spectrum showed the absorption maxima at 205, 226, and 312 nm. The IR spectrum suggested the presence of hydroxy (3417 cm−1), carbonyl (1693 cm−1), and aromatic (1608, 1515, and 1453 cm−1) groups. The 1D NMR data of 2 were similar to those of 1, except for the presence of a trans-p-coumaroyl group [δH 7.47 (2H, d, J = 8.6 Hz), 7.63 (1H, d, J = 15.9 Hz), 6.82 (2H, d, J = 8.6 Hz), 6.40 (1H, d, J = 15.9 Hz); δC 169.5, 161.3, 146.4, 131.3 (×2), 127.4, 117.0 (×2), 116.0], and the absence of a cis-p-coumaroyl group (Table 1). The structure of 2 was verified by its 2D NMR spectra. In the NOESY spectrum, the correlations between H-3/H3-25 and H3-23 indicated that H-3 was β-oriented (Figure 3). Thus, the structure of 2 was elucidated and named as 3α-O-trans-p-coumaroyloxy-2α,19α-dihydroxy-12-ursen-28-oic acid.
The molecular formula of 3 was determined as C39H54O6 by the HRESIMS at m/z 619.3986 [M + H]+ (calcd for C39H55O6, 619.3993). The UV spectrum showed the absorption maxima at 208, 228, and 312 nm. The IR spectrum suggested the presence of hydroxy (3449 cm−1), carbonyl (1695 cm−1), and aromatic (1599, 1519, and 1458 cm−1) groups. The 1D NMR data of 3 were similar to those of jacoumaric acid (11) [19,20], except for the chemical shifts of ΔδC −5.0 (C-1), −1.7 (C-2), −4.1 (C-3), −1.6 (C-4), −5.1 (C-5), −0.6 (C-23), and +3.7 (C-24), which indicated 3 might be a 3-epimer of 11. In the NOESY spectrum, the similar correlations between H-3/H3-25 and H3-23 suggested that H-3 was β-oriented. Based on the above analysis, compound 3 was determined to be 3α-O-trans-p-coumaroyloxy-2α-hydroxy-12-ursen-28-oic acid.
Compound 4 was isolated as an amorphous powder. Its HRESIMS at m/z 639.3637 [M + Na]+ exhibited the molecular formula of C39H52O6. The UV spectrum showed the absorption maxima at 208, 228, and 312 nm. The IR spectrum suggested the presence of hydroxy (3204 cm−1), carbonyl (1695 cm−1), and aromatic (1602, 1515, and 1453 cm−1) groups. The 1H and 13C NMR spectra showed a trans-p-coumaroyl group [δH 8.02 (1H, d, J = 15.9 Hz), 7.57 (2H, d, J = 8.5 Hz), 7.18 (2H, d, J = 8.5 Hz), 6.70 (1H, d, J = 15.9 Hz); δC 168.4, 161.8, 145.3, 131.1, 131.1, 126.7, 117.3, 117.3, 116.6], two olefinic bonds [δH 5.47 (1H, m), 4.84 (1H, br s), 4.79 (1H, br s); δC 154.2, 139.5, 126.2, 105.6], two oxymethines [δH 5.28 (1H, d, J = 10.8 Hz), 4.32 (1H, ddd, J = 10.8, 4.3, 3.7 Hz); δC 85.5, 66.8], and six methyls [δH 1.22 (3H, s), 1.13 (3H, d, J = 6.4 Hz), 1.09 (3H, s), 1.06 (3H, s), 1.03 (3H, s), 1.00 (3H, s); δC 29.5, 24.2, 18.7, 17.9, 17.4, 17.1] (Table 2). The above NMR data of 4 were similar to those of 3α-trans-coumaroyloxy-2α-hydroxy-12,20(30)-dien-28-ursolic acid (15) [21], except for the chemical shifts of ΔδC +5.0 (C-1), +1.7 (C-2), +4.1 (C-3), +1.3 (C-4), +5.1 (C-5), +0.7 (C-23), and −3.7 (C-24), which indicated 4 might be a 3-epimer of 15. In the NOESY spectrum, the correlations between H-2 and H3-25, and between H-3 and H-5, suggested that H-2 was β-oriented, while H-3 was α-oriented (Figure 4). Thus, the structure of 4 was elucidated and named as 3β-trans-p-coumaroyloxy-2α-hydroxy-12,20(30)-ursadien-28-oic acid.
The molecular formula of 5 was determined as C40H54O7 by the HRESIMS at m/z 647.3944 [M + H]+ (calcd for C40H55O7, 647.3942). The UV spectrum showed the absorption maxima at 208, 238, and 325 nm. The IR spectrum suggested the presence of hydroxy (3301 cm−1), carbonyl (1699 cm−1), and aromatic (1598, 1514, and 1447 cm−1) groups. Comparison of NMR data of 5 to those of 4 (Table 2) revealed similar structures, except for the presence of a methoxy group [δH 3.80 (3H, s); δC 56.4] in 5. In the HMBC spectrum, the correlation from CH3O- to C-6′ located CH3O- at C-6′ (Figure 2). Thus, the structure of 5 was elucidated and named as 3β-trans-feruloyloxy-2α-hydroxy-12,20(30)-ursadien-28-oic acid.
The molecule formula of 6 was identical to that of 5 based on the HRESIMS at m/z 647.3945 [M + H]+ (calcd for C40H55O7, 647.3942). The UV spectrum showed the absorption maxima at 208, 242, and 325 nm. The IR spectrum suggested the presence of hydroxy (3450 cm−1), carbonyl (1699 cm−1), and aromatic (1597, 1517, and 1460 cm−1) groups. The 1D NMR data of 6 were similar to those of 5, except for the chemical shifts of ΔδC +5.0 (C-1), +1.8 (C-2), +4.1 (C-3), +1.1 (C-4), +5.2 (C-5), +0.6 (C-23), and −3.7 (C-24), which indicated 6 might be a 3-epimer of 5. In the NOESY spectrum, the correlations between H-3/H3-25 and H3-23 suggested that H-3 was β-oriented. Thus, the structure of 6 was identified and named as 3α-trans-feruloyloxy-2α-hydroxy-12,20(30)-ursadien-28-oic acid.
Apart from the above six new 2α-hydroxy ursane triterpenoids (16), eleven known triterpenoids were isolated and identified as 3α-trans-feruloyloxy-2α-hydroxyurs-12-en-28-oic acid (7) [22], 3-O-trans-feruloyl euscaphic acid (8) [23], colosolic acid (9) [24], 3β-O-cis-p-coumaroyl-2α-hydroxy-urs-12-en-28-oic acid (10) [20,25], jacoumaric acid (11) [20], 3β-O-cis-feruloyl-2α-hydroxy-urs-12-en-28-oic acid (12) [20], 3β-O-trans-feruloyl-2α-hydroxy-urs-12-en-28-oic acid (13) [20], (2α,3α)-2-hydroxy-3-[(2Z)-3-(4-hydroxyphenyl)-1-oxo-2-propenyl]oxy]ursa-12,20(30)-dien-28-oic acid (14) [21], 3α-trans-coumaroyloxy-2α-hydroxy-12,20(30)-dien-28-ursolic acid (15) [21], 2α-hydroxymicromeric acid (16) [26], and 3β-cis-p-coumaroyloxy-2α-hydroxyursa-12,20(30)-dien-28-oic acid (17) [27].
All the isolated compounds were evaluated for their PTP1B inhibitory activity. As a result, compounds 46, 1013, and 15 showed PTP1B inhibition with IC50 values in the range of 10.32–48.67 μM (Table 3), while other compounds were over 50 μM. Compounds 12, 13, and 15 showed better inhibition activity with IC50 values of 16.20, 10.32, and 17.12 μM, respectively. To know more about the binding and interaction mode between PTP1B and compounds 12, 13, 15, and oleanolic acid, a molecular docking study was conducted by AutoDock Vina. The binding energies of compounds 12, 13, and 15 to PTP1B are −7.1, −7.8, and −7.5 kcal/mol, respectively. Compound 13 is slightly better than that of oleanolic acid (−7.6 kcal/mol), suggesting comparable binding affinity to PTP1B. As shown in Figure 5, these active compounds can dock into the same hydrophobic pocket and bind to the catalytic residues (Gln262, Ala217, and Tyr46) by different interactions as the positive control.

3. Experimental Section

3.1. General Experimental Procedures

UV, IR, ECD, and optical rotations were recorded on JASCO V550 UV/VIS, JASCO FTR-4600, JASCO-180, and JASCO P-2000 spectrometers (JASCO, Tokyo, Japan), respectively. HRESIMS were obtained using an Agilent 6210 LC/MSD TOF mass spectrometer (Agilent Technologies, Inc., Santa Clara, CA, USA). NMR data were measured by a Bruker AV-400 NMR spectrometer (Bruker, Fällanden, Switzerland). The preparative HPLC was carried on an Agilent 1200 HPLC (Agilent Technologies, Inc., Santa Clara, CA, USA) with a Cosmosil 5C18-MS-II (250 mm × 10 mm, 5 μm). The GF254 silica gel plates were purchased from the Yantai Institute of Industrial Chemistry, Yantai, China. Silica gel (80–100 mesh, 100–200 mesh, and 200–300 mesh; Qingdao Marine Chemical, Ltd., Qingdao, China), ODS (C18, Merck, Darmstadt, Germany), and Sephadex LH-20 (Pharmacia, Kalamazoo, MI, USA) were used for column chromatography.

3.2. Plant Material

The leaves of Diospyros digyna Jacq. were collected from Zhongshan Haizaoye Agricultural Technology Co., Ltd., Guangdong, China, in July 2018. The plant was identified by Prof. Guangxiong Zhou, College of Pharmacy, Jinan University. A voucher specimen (No. CP2018070903) was deposited in the herbarium of Jinan University.

3.3. Extraction and Isolation

The dried leaves of D. digyna (17.0 kg) were extracted with 95% ethanol at room temperature to give the crude extract (4.3 kg). Then, the crude extract was suspended into H2O, and partitioned successively with petroleum ether (PE), ethyl acetate (EA), and n-butanol, respectively. The ethyl acetate extract (803 g) was subjected to a silica gel column and eluted with CH2Cl2/CH3OH (100:0→0:100, v/v) to obtain eight main fractions (Fr. A–Fr. H). Fr C (20 g) was subjected to a silica gel column eluted with PE/EA (10:1→1:1, v/v) to afford six subfractions (Fr. C1–Fr. C6). Fr. C3 (5.0 g) was purified by Sephadex LH-20 columns (CHCl3:CH3OH = 1:1, v/v) and preparative HPLC to afford 3 (8.5 mg, CH3OH:H2O:HCOOH = 83:17:0.1, v/v/v, tR = 32.7 min), 9 (8.0 mg, CH3OH:H2O:HCOOH = 98:2:0.1, v/v/v, tR = 32.5 min), 11 (9.0 mg, CH3OH:H2O:HCOOH = 85:15:0.1, v/v/v, tR = 11.0 min), 13 (6.0 mg, CH3OH:H2O:HCOOH = 83:17:0.1, v/v/v, tR = 31.5 min), and 14 (13.0 mg, CH3OH:H2O:HCOOH = 83:17:0.1, v/v/v, tR = 39.3 min). Fr. D (15 g) was subjected to an ODS column eluted with CH3OH/H2O (100:0→0:100, v/v) to obtain seven subfractions (Fr. D1–Fr. D7). Fr. D4 (5.8 g) was purified by Sephadex LH-20 columns (CH3OH) and preparative HPLC to afford 1 (26.0 mg, CH3CN:H2O:HCOOH = 60:40:0.1, v/v/v, tR = 22.1 min), 2 (10.0 mg, CH3CN:H2O:HCOOH = 60:40:0.1, v/v/v, tR = 25.2 min), 10 (10.0 mg, CH3OH:H2O:HCOOH = 80:20:0.1, v/v/v, tR = 15.5 min), 12 (20.0 mg, CH3CN:H2O:HCOOH = 75:25:0.1, v/v/v, tR = 17.5 min), 16 (2.0 mg, CH3CN:H2O:HCOOH = 65:35:0.1, v/v/v, tR = 20.5 min), and 17 (6.5 mg, CH3CN:H2O:HCOOH = 65:35:0.1, v/v/v, tR = 22.5 min). Fr. F (10 g) was subjected to a silica gel column eluted with PE/EA (9:1→0:10, v/v) to obtain six subfractions (Fr. F1–Fr. F6). Fr. F4 (2 g) was purified by a MCI column and preparative HPLC to afford 4 (10.0 mg, CH3CN:H2O:HCOOH = 60:40:0.1, v/v/v, tR = 19.5 min), 5 (12.0 mg, CH3CN:H2O:HCOOH = 60:40:0.1, v/v/v, tR = 14.5 min), 6 (9.0 mg, CH3CN:H2O:HCOOH = 70:30:0.1, v/v/v, tR = 32.0 min), 7 (28.5 mg, CH3CN:H2O:HCOOH = 70:30:0.1, v/v/v, tR = 24.3 min), 8 (19.0 mg, CH3CN:H2O:HCOOH = 60:40:0.1, v/v/v, tR = 14.0 min), and 15 (12.0 mg, CH3OH:H2O:HCOOH = 80:20:0.1, v/v/v, tR = 12.5 min).
Compound 1: white amorphous powder; [ α ] D 25 + 16 (c 1.3, CH3OH); UV (CH3OH) λmax (log ε) 206 (3.59), 228 (3.43), 310 (3.58) nm; IR (KBr) vmax 3425, 2937, 1694, 1605, 1513, 1454, 1387, 1273, 1173, 1041, 979, 937, 839 cm−1; HRESIMS m/z: 635.3934 [M + H]+ (calcd for C39H55O7, 635.3942); 1H NMR (400 MHz, CD3OD) and 13C NMR (100 MHz, CD3OD) (Table 1 and Figures S1–S6, Supplementary Materials).
Compound 2: white amorphous powder; [ α ] D 25 + 24 (c 1.3, CH3OH); UV (CH3OH) λmax (log ε) 205 (3.62), 226 (3.46), 312 (3.64) nm; IR (KBr) vmax 3417, 2938, 1693, 1608, 1515, 1453, 1386, 1272, 1198, 1177, 1041, 960, 936, 837 cm−1; HRESIMS m/z: 635.3938 [M + H]+ (calcd for C39H55O7, 635.3942); 1H NMR (400 MHz, CD3OD) and 13C NMR (100 MHz, CD3OD) (Table 1 and Figures S10–S15, Supplementary Materials).
Compound 3: white amorphous powder; [ α ] D 25 + 35 (c 0.3, CH3OH); UV (CH3OH) λmax (log ε) 208 (3.46), 228 (3.31), 312 (3.53) nm; IR (KBr) vmax 3449, 2935, 1695, 1599, 1519, 1458, 1268, 1178, 1039, 984, 819 cm−1; HRESIMS m/z: 619.3986 [M + H]+ (calcd for C39H55O6, 619.3993); 1H NMR (400 MHz, C5D5N) and 13C NMR (100 MHz, C5D5N) (Table 1 and Figures S19–S24, Supplementary Materials).
Compound 4: white amorphous powder; [ α ] D 25 + 24 (c 1.4, CH3OH); UV (CH3OH) λmax (log ε) 208 (3.49), 228 (3.30), 312 (3.54) nm; IR (KBr) vmax 3204, 2938, 1695, 1602, 1515, 1453, 1385, 1273, 1175, 1042, 961, 938, 840 cm−1; HRESIMS m/z: 639.3637 [M + Na]+ (calcd for C39H52O6Na, 639.3656); 1H NMR (400 MHz, C5D5N) and 13C NMR (100 MHz, C5D5N) (Table 2 and Figures S28–S33, Supplementary Materials).
Compound 5: white amorphous powder; [ α ] D 25 + 94 (c 1.7, CH3OH); UV (CH3OH) λmax (log ε) 208 (3.60), 238 (3.37), 325 (3.51) nm; IR (KBr) vmax 3301, 2939, 1699, 1598, 1514, 1447, 1369, 1268, 1170, 1102, 1018, 869, 833 cm−1; HRESIMS m/z: 647.3944 [M + H]+ (calcd for C40H55O7, 647.3942); 1H NMR (400 MHz, C5D5N) and 13C NMR (100 MHz, C5D5N) (Table 2 and Figures S37–S42, Supplementary Materials).
Compound 6: white amorphous powder; [ α ] D 25 + 24 (c 1.4, CH3OH); UV (CH3OH) λmax (log ε) 208 (3.48), 242 (3.26), 325 (3.35) nm; IR (KBr) vmax 3450, 2942, 1699, 1597, 1517, 1460, 1378, 1269, 1175, 1131, 1037, 966, 942, 817 cm−1; HRESIMS m/z: 647.3945 [M + H]+ (calcd for C40H55O7, 647.3942); 1H NMR (400 MHz, C5D5N) and 13C NMR (100 MHz, C5D5N) (Table 2 and Figures S46–S51, Supplementary Materials).

3.4. PTP1B Inhibition Assay

The inhibitory activity of the isolated compounds against PTP1B (Abcam, Cambridge, UK, human recombinant) was assayed according to the method reported previously [28]. The reagent p-nitrophenyl phosphate (pNPP) was used as the substrate, and oleanolic acid was used as the positive control. In brief, a 100 μL assay mixture containing 1 μg/mL PTP1B, samples, 4 mM pNPP, 55 mM NaCl, 2.2 mM DTT, 1.1 mM EDTA, and 1 mM BSA in 11 mM Tris-HCl, pH 7.5, was incubated at 37 °C for 30 min in a 96-well plate. The absorbance of 405 nm was measured by a microplate reader. Data were analyzed by GraphPad Prism v.10.2.0 software. All data were obtained in triplicate and presented as means ± SD.

3.5. Molecular Docking Analysis

The crystal structure of PTP1B (PDB ID: 8U1E) was obtained from the RCSB Protein Data Bank database. The receptor was prepared by PyMOL 2.5.0 and deposited as a .pdb format. The 3D structures of the ligands were energy optimized by ChemDraw 3D 18.0 and deposited as a mol2 format. The ligands and receptors for molecular docking analysis were conducted by AutoDock Vina 1.2.2 [29]. The grid box parameters (X-center = 0.405, Y-center = 12.132, Z-center = 25.000; x-dimension = 46, y-dimension = 60, z-dimension = 48) were set to cover the binding pocket in the receptor. The docking calculation results were analyzed by PyMOL 2.5.0.

4. Conclusions

In summary, six new 2α-hydroxy ursane triterpenoids (16), along with eleven known ursane triterpenoids (717), were isolated from the leaves of D. digyna. Compounds 46, 1013, and 15 showed PTP1B inhibitory activity. Notably, compound 13 demonstrated PTP1B inhibition comparable to that of oleanolic acid (positive control). The structure–activity relationships of these triterpenoids were briefly summarized. Compound 13 showed stronger PTP1B inhibitory activity than those of 1012, indicating that the trans-feruloyl group at C-3 strengthens the activity. For compounds 3, 7, 11, and 13, the α-orientation of the substituents at C-3 weakened the activity. Moreover, the molecular docking study further confirmed the binding affinity between compound 13 and PTP1B. The naturally occurring PTP1B inhibitors might reveal the potential utilization of D. digyn in the treatment of T2DM, and their action mechanisms deserve further investigation.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules29071640/s1, Figures S1–S54: 1D, 2D NMR, HRESIMS, IR, and UV spectra of compounds 16.

Author Contributions

L.H. and Z.W.: investigation, methodology, data curation, writing—original draft; F.W., S.W. and D.W.: resources, investigation, validation; M.G.: resources, formal analysis; X.Z., M.S. and H.L.: conceptualization, writing—review and editing, funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the National Natural Science Foundation of China (Nos. 82073712, 82304330, U1801287), Science and Technology Key Project of Guangdong Province (Nos. 2020B1111110004, 2022A1515011850), Science and Technology Planning Project of Guangzhou City (Nos. 20212210005, 202102070001), China Postdoctoral Science Foundation (2023M731325), Science and Technology Project of Guangdong Provincial Medical Products Administration (Nos. 2022YDZ02, 2023ZDZ01), and the High-performance Super Computing Platform of Jinan University.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in this article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Galicia-Garcia, U.; Benito-Vicente, A.; Jebari, S.; Larrea-Sebal, A.; Siddiqi, H.; Uribe, K.B.; Ostolaza, H.; Martín, C. Pathophysiology of type 2 diabetes mellitus. Int. J. Mol. Sci. 2020, 21, 6275. [Google Scholar] [CrossRef] [PubMed]
  2. Burns, C.; Sirisena, I. Type 2 diabetes-etiology, epidemiology, pathogenesis, treatment. In Metabolic Syndrome: A Comprehensive Textbook; Ahima, R.S., Ed.; Springer International Publishing: Cham, Switzerland, 2014; pp. 1–19. [Google Scholar]
  3. Salmeen, A.; Andersen, J.N.; Myers, M.P.; Tonks, N.K.; Barford, D. Molecular basis for the dephosphorylation of the activation segment of the insulin receptor by protein tyrosine phosphatase 1B. Mol. Cell 2000, 6, 1401–1412. [Google Scholar] [CrossRef] [PubMed]
  4. Dubé, N.; Tremblay, M.L. Involvement of the small protein tyrosine phosphatases TC-PTP and PTP1B in signal transduction and diseases: From diabetes, obesity to cell cycle, and cancer. Biochim. Biophys. Acta 2005, 1754, 108–117. [Google Scholar] [CrossRef] [PubMed]
  5. Elchebly, M.; Payette, P.; Michaliszyn, E.; Cromlish, W.; Collins, S.; Loy, A.L.; Normandin, D.; Cheng, A.; Himms-Hagen, J.; Chan, C.C.; et al. Increased insulin sensitivity and obesity resistance in mice lacking the protein tyrosine phosphatase-1B gene. Science 1999, 283, 1544–1548. [Google Scholar] [CrossRef] [PubMed]
  6. Jensen-Cody, S.; Coyne, E.S.; Ding, X.; Sebin, A.; Vogel, J.; Goldstein, J.; Rosahl, T.W.; Zhou, H.H.; Jacobs, H.; Champy, M.F.; et al. Loss of low-molecular-weight protein tyrosine phosphatase shows limited improvement in glucose tolerance but causes mild cardiac hypertrophy in mice. Am. J. Physiol. Endocrinol. Metab. 2022, 322, E517–E527. [Google Scholar] [CrossRef] [PubMed]
  7. Zhang, Z.; Shang, Z.P.; Jiang, Y.; Qu, Z.X.; Yang, R.Y.; Zhang, J.; Lin, Y.X.; Zhao, F. Selective inhibition of PTP1B by new anthraquinone glycosides from Knoxia valerianoides. J. Nat. Prod. 2022, 85, 2836–2844. [Google Scholar] [CrossRef]
  8. Zhao, J.F.; Li, L.H.; Guo, X.J.; Zhang, H.X.; Tang, L.L.; Ding, C.H.; Liu, W.S. Identification of natural product inhibitors of PTP1B based on high-throughput virtual screening strategy: In silico, in vitro and in vivo studies. Int. J. Biol. Macromol. 2023, 243, 125292. [Google Scholar] [CrossRef]
  9. Liu, R.; Mathieu, C.; Berthelet, J.; Zhang, W.; Dupret, J.M.; Rodrigues Lima, F. Human protein tyrosine phosphatase 1B (PTP1B): From structure to clinical inhibitor perspectives. Int. J. Mol. Sci. 2022, 23, 7027. [Google Scholar] [CrossRef] [PubMed]
  10. Newman, D.J.; Cragg, G.M. Natural products as sources of new drugs over the nearly four decades from 01/1981 to 09/2019. J. Nat. Prod. 2020, 83, 770–803. [Google Scholar] [CrossRef]
  11. Jiang, C.S.; Liang, L.F.; Guo, Y.W. Natural products possessing protein tyrosine phosphatase 1B (PTP1B) inhibitory activity found in the last decades. Acta Pharmacol. Sin. 2012, 33, 1217–1245. [Google Scholar] [CrossRef]
  12. Liu, Z.; Gao, H.; Zhao, Z.; Huang, M.; Wang, S.; Zhan, J. Status of research on natural protein tyrosine phosphatase 1B inhibitors as potential antidiabetic agents: Update. Biomed. Pharmacother. 2023, 157, 113990. [Google Scholar] [CrossRef]
  13. Ramírez-Briones, E.; Rodríguez Macías, R.; Casarrubias Castillo, K.; del Río, R.E.; Martínez-Gallardo, N.; Tiessen, A.; Ordaz-Ortiz, J.; Cervantes-Hernández, F.; Délano-Frier, J.P.; Zañudo-Hernández, J. Fruits of wild and semi-domesticated Diospyros tree species have contrasting phenological, metabolic, and antioxidant activity profiles. J. Sci. Food Agric. 2019, 99, 6020–6031. [Google Scholar] [CrossRef] [PubMed]
  14. Yahia, E.M.; Gutierrez-Orozco, F.; Leon, C.A.D. Phytochemical and antioxidant characterization of the fruit of black sapote (Diospyros digyna Jacq.). Food Res. Int. 2011, 44, 2210–2216. [Google Scholar] [CrossRef]
  15. Rauf, A.; Uddin, G.; Patel, S.; Khan, A.; Halim, S.A.; Bawazeer, S.; Ahmad, K.; Muhammad, N.; Mubarak, M.S. Diospyros, an under-utilized, multi-purpose plant genus: A review. Biomed. Pharmacother. 2017, 91, 714–730. [Google Scholar] [CrossRef] [PubMed]
  16. Thuong, P.T.; Lee, C.H.; Dao, T.T.; Nguyen, P.H.; Kim, W.G.; Lee, S.J.; Oh, W.K. Triterpenoids from the leaves of Diospyros kaki (Persimmon) and their inhibitory effects on protein tyrosine phosphatase 1B. J. Nat. Prod. 2008, 71, 1775–1778. [Google Scholar] [CrossRef]
  17. Ramírez-Briones, E.; Rodríguez-Macías, R.; Salcedo-Pérez, E.; Ramírez-Chávez, E.; Molina-Torres, J.; Tiessen, A.; Ordaz-Ortiz, J.; Martínez-Gallardo, N.; Délano-Frier, J.P.; Zañudo-Hernández, J. Seasonal changes in the metabolic profiles and biological activity in leaves of Diospyros digyna and D. rekoi “Zapote” trees. Plants 2019, 8, 449. [Google Scholar] [CrossRef] [PubMed]
  18. Taniguchi, S.; Imayoshi, Y.; Kobayashi, E.; Takamatsu, Y.; Ito, H.; Hatano, T.; Sakagami, H.; Tokuda, H.; Nishino, H.; Sugita, D.; et al. Production of bioactive triterpenes by Eriobotrya japonica calli. Phytochemistry 2002, 59, 315–323. [Google Scholar] [CrossRef]
  19. Ogura, M.; Cordell, G.A.; Farnsworth, N.R. Jacoumaric acid, a new triterpene ester from Jacaranda caucana. Phytochemistry 1977, 16, 286–287. [Google Scholar] [CrossRef]
  20. Häberlein, H.; Tschiersch, K.P. Triterpenoids and flavonoids from Leptospermum scoparium. Phytochemistry 1994, 35, 765–768. [Google Scholar] [CrossRef]
  21. Cai, Y.H.; Guo, Y.; Li, Z.; Wu, D.; Li, X.; Zhang, H.; Yang, J.; Lu, H.; Sun, Z.; Luo, H.B.; et al. Discovery and modelling studies of natural ingredients from Gaultheria yunnanensis (FRANCH.) against phosphodiesterase-4. Eur. J. Med. Chem. 2016, 114, 134–140. [Google Scholar] [CrossRef]
  22. Ito, H.; Kobayashi, E.; Li, S.H.; Hatano, T.; Sugita, D.; Kubo, N.; Shimura, S.; Itoh, Y.; Yoshida, T. Megastigmane glycosides and an acylated triterpenoid from Eriobotrya japonica. J. Nat. Prod. 2001, 64, 737–740. [Google Scholar] [CrossRef] [PubMed]
  23. Shimizu, M.; Eumitsu, N.; Shirota, M.; Matsumoto, K.; Tezuka, Y. A new triterpene ester from Eriobotrya japonica. Chem. Pharm. Bull. 1996, 44, 2181–2182. [Google Scholar] [CrossRef]
  24. Glen, A.T.; Lawrie, W.; McLean, J.; Younes, M.E.G. Triterpenoid constituents of rose-bay willow-herb. J. Chem. Soc. C 1967, 510–515. [Google Scholar] [CrossRef]
  25. Chao, I.C.; Chen, Y.; Gao, M.H.; Lin, L.G.; Zhang, X.Q.; Ye, W.C.; Zhang, Q.W. Simultaneous determination of α-glucosidase inhibitory triterpenoids in Psidium guajava using HPLC–DAD–ELSD and pressurized liquid extraction. Molecules 2020, 25, 1278. [Google Scholar] [CrossRef] [PubMed]
  26. Wang, C.; Wu, P.; Tian, S.; Xue, J.; Xu, L.; Li, H.; Wei, X. Bioactive pentacyclic triterpenoids from the leaves of Cleistocalyx operculatus. J. Nat. Prod. 2016, 79, 2912–2923. [Google Scholar] [CrossRef] [PubMed]
  27. Lee, C.K. Ursane triterpenoids from leaves of Melaleuca leucadendron. Phytochemistry 1998, 49, 1119–1122. [Google Scholar] [CrossRef]
  28. Zhou, J.; Sun, N.; Zhang, H.; Zheng, G.; Liu, J.; Yao, G. Rhodomollacetals A–C, PTP1B inhibitory diterpenoids with a 2,3:5,6-di-seco-grayanane skeleton from the leaves of Rhododendron molle. Org. Lett. 2017, 19, 5352–5355. [Google Scholar] [CrossRef]
  29. Trott, O.; Olson, A.J. AutoDock Vina: Improving the speed and accuracy of docking with a new scoring function, efficient optimization, and multithreading. J. Comput. Chem. 2010, 31, 455–461. [Google Scholar] [CrossRef]
Figure 1. Chemical structures of 117.
Figure 1. Chemical structures of 117.
Molecules 29 01640 g001
Figure 2. Key 1H–1H COSY and HMBC correlations of 1, 4, and 5.
Figure 2. Key 1H–1H COSY and HMBC correlations of 1, 4, and 5.
Molecules 29 01640 g002
Figure 3. Key NOESY correlations of 1 and 2.
Figure 3. Key NOESY correlations of 1 and 2.
Molecules 29 01640 g003
Figure 4. Key NOESY correlations of 4 and 5.
Figure 4. Key NOESY correlations of 4 and 5.
Molecules 29 01640 g004
Figure 5. Three-dimensional ligand interaction diagrams of oleanolic acid (A), 12 (B), 13 (C), and 15 (D) at the active site of PTP1B enzyme (blue dashed lines indicate hydrogen bond, gray dashed lines indicate hydrophobic interaction).
Figure 5. Three-dimensional ligand interaction diagrams of oleanolic acid (A), 12 (B), 13 (C), and 15 (D) at the active site of PTP1B enzyme (blue dashed lines indicate hydrogen bond, gray dashed lines indicate hydrophobic interaction).
Molecules 29 01640 g005
Table 1. 1H (400 MHz) and 13C NMR (100 MHz) data of 13 (δ in ppm, J in Hz) a.
Table 1. 1H (400 MHz) and 13C NMR (100 MHz) data of 13 (δ in ppm, J in Hz) a.
No. 1 b2 b3 c
δHδCδHδCδHδC
1β1.6543.21.7143.52.0544.1
α1.37 1.38 1.86, m
2 4.11, dt (11.0, 4.1)66.24.12, dt (10.4, 4.1)66.24.50, dt (11.4, 4.3)65.2
3 5.00, d (4.1)81.45.03, d (4.1)81.75.62, d (4.3)81.4
4 39.6 39.8 39.2
5 1.19, m51.11.3051.31.56, m50.9
6α1.4719.31.5019.31.49, m18.8
β1.47 1.45 1.34
7α1.59, m34.11.6034.21.67, m33.8
β1.34 1.35 1.41
8 41.4 41.4 40.6
9 1.8248.61.94, m48.82.0048.6
10 39.6 39.7 39.2
11α2.04, m24.82.06, m24.92.1224.2
β1.45 1.36 2.02
12 5.32, m129.45.34, m129.45.51, m125.9
13 140.2 140.3 139.9
14 42.8 42.8 43.0
15β1.8329.71.8329.72.37, m29.1
α1.03 1.02 1.21, m
16α2.60, m26.72.61, m26.72.1125.4
β1.53, m 1.54 1.96
17 48.8 49.0 48.5
18 2.54, s55.22.54, s55.22.67, m54.0
19 73.8 73.81.4539.9
20 1.3143.31.3343.21.0439.9
21α1.7327.41.7327.41.4031.6
β1.27 1.23 1.48
22β1.7439.21.7539.22.0038.0
α1.64 1.65 2.00
23 0.91, s28.70.92, s28.71.13, s28.9
24 1.00, s22.41.01, s22.30.96, s21.9
25 1.04, s17.01.07, s17.01.00, s17.1
26 0.81, s17.70.84, s17.71.08, s18.0
27 1.36, s25.11.44, s25.11.25, s24.5
28 182.4 182.4 180.6
29 1.23, s27.21.23, s27.21.0022.5
30 0.95, d (6.7)16.80.96, d (6.6)16.81.0018.0
1’ 168.6 169.5 168.4
2’ 5.87, d (13.0)117.56.40, d (15.9)116.06.78, d (15.9)116.6
3’ 6.88, d (13.0)144.87.63, d (15.9)146.48.03, d (15.9)145.5
4’ 127.9 127.4 126.7
5’ 7.66, d (8.7)133.87.47, d (8.6)131.37.52, d (8.6)131.1
6’ 6.75, d (8.7)116.06.82, d (8.6)117.07.15, d (8.6)117.2
7’ 160.0 161.3 161.8
8’ 6.75, d (8.7)116.06.82, d (8.6)117.07.15, d (8.6)117.2
9’ 7.66, d (8.7)133.87.47, d (8.6)131.37.52, d (8.6)131.1
a Overlapped signals were reported without designating multiplicity. b NMR data were recorded in CD3OD. c NMR data were recorded in C5D5N.
Table 2. 1H (400 MHz) and 13C NMR (100 MHz) data of 46 (C5D5N, δ in ppm, J in Hz) a.
Table 2. 1H (400 MHz) and 13C NMR (100 MHz) data of 46 (C5D5N, δ in ppm, J in Hz) a.
No. 456
δHδCδHδCδHδC
1β2.3349.02.3349.12.0344.1
α1.42, m 1.41, m 1.86, m
2 4.32, ddd (10.8, 4.3, 3.7)66.84.32, ddd (10.8, 4.3, 3.7)66.94.49, dt (10.8, 4.5)65.1
3 5.28, d (10.8)85.55.29, d (10.8)85.55.64, br s81.4
4 40.4 40.3 39.2
5 1.10, m56.01.11, m56.11.5350.9
6α1.5119.11.5319.11.4718.7
β1.37 1.38 1.33
7α1.5733.71.5633.81.62, m33.7
β1.35 1.37 1.39
8 40.3 40.5 40.6
9 1.74, m48.41.76, m48.51.9548.5
10 38.7 38.8 39.2
11α1.9824.21.9924.22.0624.1
β1.98 1.27 1.27
12 5.47, m126.25.48, m126.25.49, m126.3
13 139.5 139.5 139.5
14 43.1 43.1 43.0
15β2.3129.12.3229.12.3329.0
α1.24, m 1.25, m 1.22, m
16α2.3025.32.3225.42.3025.3
β2.12 2.10 2.09
17 48.7 48.8 48.7
18 2.77, d (11.8)56.02.78, d (11.8)56.12.76, d (11.6)56.0
19 2.48, m38.22.48, m38.22.44, m38.2
20 154.2 154.3 154.2
21α2.2733.22.3033.22.2833.2
β2.43, m 2.45, m 2.40
22β2.1440.12.1540.22.12, m40.1
α2.03 2.02 2.02
23 1.09, s29.51.10, s29.51.14, s28.9
24 1.06, s18.71.05, s18.80.96, s22.5
25 1.00, s17.41.01, s17.40.98, s17.1
26 1.03, s17.91.04, s17.91.05, s17.9
27 1.22, s24.21.23, s24.31.17, s24.2
28 179.8 179.9 179.8
29 1.13, d (6.4)17.11.14, d (6.1)17.11.11, d (6.1)17.1
30a4.84, br s105.64.84, br s105.64.83, br s105.6
b4.79, br s 4.79, br s 4.79, br s
1′ 168.4 168.4 168.4
2′ 6.70, d (15.9)116.66.75, d (15.8)116.76.89, d (15.9)116.7
3′ 8.02, d (15.9)145.38.04, d (15.8)145.68.07, d (15.9)146.0
4′ 126.7 127.2 127.1
5′ 7.57, d (8.5)131.17.31, m112.07.24, m111.8
6′ 7.18, d (8.5)117.3 149.5 149.4
7′ 161.8 151.5 151.5
8′ 7.18, d (8.5)117.37.23, m117.37.20, m117.2
9′ 7.57, d (8.5)131.17.23, m124.17.20, m124.5
Ome 3.80, s56.43.72, s56.3
a Overlapped signals were reported without designating multiplicity.
Table 3. PTP1B inhibitory activity.
Table 3. PTP1B inhibitory activity.
CompoundsIC50 (μM)CompoundsIC50 (μM)
448.67 ± 5.171126.20 ± 1.31
523.74 ± 0.731216.20 ± 0.57
634.01 ± 4.881310.32 ± 1.21
1019.15 ± 0.221517.12 ± 1.67
Oleanolic acid a10.19 ± 0.12
a Positive control.
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

Huang, L.; Wang, Z.; Wang, F.; Wang, S.; Wang, D.; Gao, M.; Li, H.; Song, M.; Zhang, X. Triterpenoids from the Leaves of Diospyros digyna and Their PTP1B Inhibitory Activity. Molecules 2024, 29, 1640. https://doi.org/10.3390/molecules29071640

AMA Style

Huang L, Wang Z, Wang F, Wang S, Wang D, Gao M, Li H, Song M, Zhang X. Triterpenoids from the Leaves of Diospyros digyna and Their PTP1B Inhibitory Activity. Molecules. 2024; 29(7):1640. https://doi.org/10.3390/molecules29071640

Chicago/Turabian Style

Huang, Lan, Ziqi Wang, Fangxin Wang, Song Wang, Dezhi Wang, Meihua Gao, Hua Li, Min Song, and Xiaoqi Zhang. 2024. "Triterpenoids from the Leaves of Diospyros digyna and Their PTP1B Inhibitory Activity" Molecules 29, no. 7: 1640. https://doi.org/10.3390/molecules29071640

Article Metrics

Back to TopTop