ABSTRACT
The chemical analysis of the melliferous plant Viguiera dentata (Asteraceae) yielded cycloartanes 1–9 (including the new compounds 7–9), ent‐kaurenes (10–15), diversifolin (16) and other constituents. The structure of 9 was confirmed by X‐ray analysis. To evaluate the insecticidal potential of its constituents, in silico pesticide‐likeness calculations for structures 1–16 were performed indicating no violations of the Tice rules. Tests for activity against acetylcholinesterase revealed that only cycloartanes 1 (78.85 µM) and 6 (53.54 µM) inhibited the enzyme. Molecular docking analysis showed interactions between compounds 1, 10, and 13, with Y337, a key amino acid in the catalytic site. A bioassay against Spodoptera frugiperda revealed that compounds 1, 2, 9, 10, and 13 displayed activity (50% lethal concentration for larval mortality [LC50] 51.61, 84.56, 99.66, 24.69, and 62.40 ppm, respectively; reference: betulinic acid LC50 94.25 ppm). Thus, specific cycloartanes and ent‐kaurenes were identified as insecticidal compounds of V. dentata against S. frugiperda.
Keywords: acetylcholinesterase, in silico pesticide‐like properties, insecticidal activity, natural products, terpenes
The chemical analysis of aerial parts of the melliferous Viguiera dentata led to the isolation of 20 substances, including the new cycloartanes 24‐epi‐argentatin C, 7β‐hydroxy‐24‐epi‐argentatin C, 7β‐hydroxy‐argentatin B. In‐combo results indicated that specific argentatins and ent‐kaurenoic acids have the potential to control Spodoptera frugiperda
.
1. Introduction
Viguiera (Asteraceae, Heliantheae, Helianthinae) is an American genus initially described by Blake in 1918, with 141 species [1], and revised by Schilling and Panero in 2011 [2] to have most of its species segregated into ten genera [3]. This last classification limited the genus to nine species [4]. Previous chemical studies on the melliferous V. dentata (common name: tajonal) have led to the isolation mainly of ent‐kaurenoids and cycloartane derivatives, together with diversifolin, spathulenol, and manool [5, 6, 7, 8]. The composition of its essential oil and biological activities have also been described [9], including the volatile components of the nectariferous flowers, whose honey is much appreciated [10, 11]. The present study describes the chemical composition of a population of V. dentata (Cav.) Spreng. Nine cycloartane derivatives (1–9), including the previously unreported (7–9), six ent‐kauranes (10–15), one sesquiterpene lactone (16), verbenol, caryophyllene oxide, and two phytosterols, were isolated. Cycloartane‐type compounds exhibit a wide range of pharmacological properties, including anti‐tumor, anti‐osteoporosis, anti‐VIH, anti‐parasite, and anti‐tuberculosis activities [12], as well as insecticidal properties [13]. Similarly, ent‐kaurane derivatives possess anti‐inflammatory, anti‐tumor, anti‐bacterial [14], insecticidal, and antifeedant properties [15].
In silico evaluation of the pesticide‐likeness properties of natural products is a valuable tool to assess their biological potential [16], and most cycloartanes and ent‐kauranes isolated from V. dentata exhibited pesticide potential, as they did not violate the Tice rules. Even though this enzyme (acetylcholinesterase [AChE]) is a major target in the development of new insecticides [17], only a few studies have examined its AChE inhibitory activities. Previous docking studies have indicated that some ent‐kauranes have insecticidal effects by inhibiting AChE [18]. AChE inhibitors bind to the enzyme and interfere with the breakdown of acetylcholine, disrupting neurotransmission [17]. None of the isolated compounds has been tested as AChE inhibitors, except ent‐kaurenoic acid (10) [19, 20]. The AChE inhibitory activity of the isolated compounds was evaluated, and their interaction with the enzyme was determined using molecular docking techniques. Additionally, four cycloartanes (1, 2, 6, and 9) and two ent‐kaurene acids (10 and 13) were tested for insecticidal activity against Spodoptera frugiperda, a major maize pest.
2. Results and Discussion
Chemical analysis of V. dentata led to the identification of argentatin B (1) [21, 22], argentatin D (2) [23], cycloartanones (3–6) [5], 24‐epi‐argentatin C (7), 7β‐hydroxy‐24‐epi‐argentatin C (8), 7β‐hydroxyargentatin B (9), ent‐kaurenoic acid (10) [24], angeloylgrandifloric acid (11) [25], grandiflorenic acid (12) [21], 12α‐hydroxy‐ent‐kaur‐9(11),16‐dien‐19‐oic acid (13) [26], ent‐12‐oxo‐kaura‐9(11),16‐dien‐19‐oic acid (14) [27], 15α,16α‐epoxy‐ent‐17‐hydroxykauran‐19‐oic‐acid (15) [28], and diversifolin (16) [6] (Figure 1). Additionally, β‐sitosterol, stigmasterol, verbenol, and caryophyllene oxide were characterized. The structures of these compounds were elucidated using spectroscopic and spectrometric methods, with known compounds compared to published data and authentic samples.
FIGURE 1.
Chemical structures of compounds 1–17.
2.1. Structural Identification of New Compounds
Compound 7 had the molecular formula C30H50O4 determined from high‐resolution direct analysis in real‐time mass spectrometry (HRDARTMS) and carbon‐13 nuclear magnetic resonance (13C NMR). Its infrared (IR) spectrum exhibited absorption bands for hydroxy (3402 cm−1) and carbonyl (1701 cm−1) groups. The NMR spectra (Tables 1 and 2) were similar to those of known compounds 1–5, featuring six tertiary and one secondary methyl group, and the characteristic protons of a tetra‐substituted cyclopropyl methylene group, indicating a cycloartane skeleton. The proton NMR (1H NMR) spectrum exhibited the resonances of two oxygenated methine protons at δH 4.49 (dt 5.1, 7.8 Hz) and δH 3.39 (dd 10.5, 1.8 Hz), assigned to H‐16 and H‐24, respectively, by their correlation spectroscopy (COSY) and heteronuclear multiple bond correlations (HMBCs). The 13C NMR spectrum exhibited 30 carbon atoms, including a carbonyl group, the oxymethine carbons C‐16 (δC 72.7) and C‐24 (δC 80.4), and a tertiary carbon‐bearing oxygen atom at C‐25 (δC 73.0). HMBC correlations of CH3‐28 (δH 1.06 s), CH3‐29 (δH 1.11 s), H2‐2 (δH 2.72, td 13.8, 6.3 Hz, H‐2β and 2.31, brd 13.8, H‐2α), and H2‐1 (δH 1.86 m, H‐1α and 1.56 m, H‐1β) with the carbonyl group at δC 216.6 defined as a ketone group at C‐3. Compound 7 had the same substitution pattern as argentatin C (17) [23], and an identical CD spectrum, with a negative Cotton effect at λ 299 nm. However, its melting point, optical rotation, and thin layer chromatography (TLC) retention factor differed from those of authentic sample 17 [23]. The resonances of the side chain atoms C‐20 to C‐24 shifted downfield Δδ by 4.9, 0.7, 2.7, 2.4, and 5.2 ppm, respectively, compared to those of 17 (Table 2). Since the connectivity and stereochemistry of the cycloartane moiety in these compounds were identical, compound 7 should be the 24S epimer of argentatin C. These results are similar to reports of several 24S,25‐dihydroxycycloartane derivatives [29, 30]. Spectroscopic data for compound 17 are provided in Tables 1 and 2, as they are not available in the literature.
TABLE 1.
Proton nuclear magnetic resonance (1H NMR) spectroscopic data (CDCl3) for compounds 7–9 and 17 (δH, J in Hz).
Position | 7 [a] | 8[ b , d ] | 9[ c , d ] | 17 [a] |
---|---|---|---|---|
1a | 1.86, m | 1.82, m | 1.82, td (13.6, 4.4) | 1.87, m |
1b | 1.56, m | 1,65, m | 1.65, m | 1.58, m |
2β | 2.72, td (13.8, 6.3) | 2.78, td (14.0, 6.3) | 2.76, td (14.0, 6.8) | 2.71, td (13.8, 6.3) |
2α | 2.31, brd (13.8) | 2.27, dt (14.0, 2.1) | 2.30, ddd (14.0, 4.4, 2.8) | 2.31, ddd (13.8, 4.2, 3.0) |
5α | 1.72, m | 1.93, dd (13.3, 4.2) | 1.91, dd (13.2, 4.0) | 1.71, m |
6α | 1.61, m | 1.71, m | 1.68, m | 1.63, m |
6β | 1.17–1.11, m | 1.18, m | 1.20, m | 1.17–1.11, m |
7α | 1.40, m | 3.56 ddd (10.5, 9.1, 3.5) | 3.62, ddd (12.6, 9.2, 4.0) | 1.0–1.3, m |
7b | 1.17–1.11, m | 1.17‐1.11, m | ||
8β | 1.70, m | 1.78, d (9.1) | 1.76, d (9.2) | 1.70, m |
11a | 2.04, m | 1.98, ddd (15.4, 9.8, 5.6) | 1.94, m | 2.09, m |
11b | 1.17–1.11, m | 1.36, ddd (15.4, 10.5, 4.9) | 1.37, m | 1.2‐1.1, m |
12 | 1.67, m | 1.82, m | 1.69, m | 1.68, m |
15α | 2.03, dd (13.2, 7.8) | 2.26, dd (14.0, 5.6) | 2.13, m | 2.07, m |
15β | 1.39, dd 13.2, 5.1) | 1.67, m | 1.68, m | 1.37, m |
16 | 4.49, dt (5.1, 7.8) | 4.42, dt (5.6, 8.4) | 4.65, q (7.6) | 4.50, dt (5.1, 7.8) |
17 | 1.70, m | 1.65, m | 1.59, m | 1.70, m |
18 | 1.20, s | 1.24, s | 1.20, s | 1.20, s |
19a | 0.82, d (4.2) | 0.89, d (4.9) | under CH3 21 | 0.83, d (4.5) |
19b | 0.59, d (4.2) | 0.66, d (4.9) | 0.60, d (4.4) | 0.60, d (4.5) |
20 | 1.85, m | 1.85, m | 2.08, m | 1.91, m |
21 | 0.96, d (6.0) | 1.00, d (7.0) | 0.95, d (6.4) | 0.95, d (6.6) |
22a | 1.81, m | 1.93, m | 1.70, m | 1.74, m |
22b | 1.17–1.11, m | 1.05, m | 1.40, m | 1.60, m |
23 | 1.70, m | 1.70, m | 2.03, m | 1.61, m |
23b | 1.32, m | 1.21, m | 1.62, m | 1.41, m |
24 | 3.39, dd (10.5, 1.8) | 3.26, dd (11.2, 2.1) | 3.53, dd (12.4, 2.0) | 3.59, dd (11.3, 2.7) |
26 | 1.17, s | 1.12, s | 1.13, s | 1.26, s |
27 | 1.22, s | 1.16, s | 1.08, s | 1.17, s |
28 | 1.06, s | 1.06, s | 1.03, s | 1.06, s |
29 | 1.11, s | 1.14, s | 1.11, s | 1.11, s |
30 | 0.91, s | 0.98, s | 0.95, s | 0.91, s |
300 MHz.
700 MHz.
400 MHz.
CD3OD.
TABLE 2.
Carbon‐13 nuclear magnetic resonance (13C NMR) spectroscopic data (CDCl3) for compounds 7–9 and 17.
Position, type | 7 [a] | 8 [ b , d ] | 9 [ c , d ] | 17 [a] |
---|---|---|---|---|
1, CH2 | 33.4 | 32.5 | 32.4 | 33.4 |
2, CH2 | 37.4 | 36.7 | 36.7 | 37.4 |
3, C | 216.6 | 217.0 | 216.9 | 216.4 |
4, C | 50.2 | 49.5 | 49.4 | 50.2 |
5, CH | 48.4 | 47.0 | 46.8 | 48.4 |
6, CH2 | 21.4 | 30.9 | 30.8 | 21.4 |
7, CH2 | 25.9 | 69.7 e | 69.5 [e] | 25.9 |
8, CH | 47.8 | 54.2 | 53.6 | 47.9 |
9, C | 20.9 | 20.6 | 20.9 | |
10, C | 26.0 | 26.4 | 26.3 | 26.1 |
11, CH2 | 26.4 | 26.3 | 26.2 | 26.4 |
12, CH2 | 32.5 | 32.3 | 32.3 | 32.6 |
13, C | 45.3 | 45.5 | 45.9 | 45.4 |
14, C | 46.7 | 46.0 | 45.1 | 46.7 |
15, CH2 | 47.7 | 48.9 | 45.6 | 47.8 |
16, CH | 72.7 | 72.0 | 74.6 | 72.9 |
17, CH | 57.0 | 55.9 | 56.4 | 56.9 |
18, CH3 | 19.0 | 17.4 | 17.1 | 19.0 |
19, CH2 | 29.8 | 28.1 | 27.1 | 29.8 |
20, CH | 31.5 | 30.5 (30.8)[ f ] | 28.9 | 26.6 |
21, CH3 | 18.4 | 17.5 (18.2)[ f ] | 17.8 | 17.7 |
22, CH2 | 33.9 | 33.9 (33.9)[ f ] | 35.6 | 31.2 |
23, CH2 | 28.6 | 27.5 (28.8)[ f ] | 22.4 | 26.2 |
24, CH | 80.4 | 79.4 (80.5)[ f ] | 82.5 | 75.2 |
25, C | 73.0 | 72.5 | 73.0 | 73.0 |
26, CH3 | 26.6 | 24.3 | 25.1 | 26.7 |
27, CH3 | 23.1 | 23.5 | 23.4 | 22.1 |
28, CH3 | 22.1 | 21.2 | 21.2 | 23.0 |
29, CH3 | 20.8 | 19.8 | 19.8 | 20.8 |
30, CH3 | 20.0 | 18.5 | 20.3 | 19.9 |
75 MHz.
175 MHz.
100 MHz.
CD3OD.
CH.
CDCl3.
Compound 8 had the molecular formula C30H50O5 based on its HRDARTMS and 13C NMR data. Its IR spectrum exhibited absorption bands for hydroxy (3351 cm−1) and carbonyl (1702 cm−1) groups. The NMR spectra showed the resonances of a cycloartane derivative, similar to those of 7 (Tables 1 and 2), with an additional proton geminal to a hydroxy group at δH 3.56 (ddd 10.5, 9.1, 3.5; δC 69.7). Correlations of this resonance in the COSY experiment, with H2‐6 (δH 1.71 m, 6α and 1.18 m, 6β) and H‐8 (δH 1.78, d 9.1 Hz), and in the HMBC spectrum, with C‐5 (δC 47.0), C‐8 (δC 54.2), and C‐14 (δC 46.0), allowed to locate the hydroxy group at C‐7. Nuclear overhauser effect spectroscopy (NOESY) correlations between H‐7, H‐5 (δH 1.93, dd 13.3, 4.2 Hz), H‐6α (δH 1.71 m), CH3‐30 (δH 0.98 s), and H‐15α (δH 2.26, dd 14.0, 5.6 Hz) defined the β‐orientation of the hydroxy group at C‐7. Compound 8 also likely had a 24S configuration, as the resonances of the side chain were very similar to those of compound 7 (Table 2), and its electronic circular dichroism (ECD) spectrum showed the same pattern as compounds 7–9, with a negative Cotton effect at λ 300 nm.
Compound 9 exhibited the molecular formula C30H48O4, confirmed by HRDARTMS and 13C NMR analyses. The IR spectrum showed the hydroxy (3432 cm−1) and carbonyl (1701 cm−1) groups. The 1H NMR spectrum (Table 1), in addition to the characteristic resonances of a cycloartane‐type compound, revealed oxymethine protons of H‐16 (δH 4.65, q, 7.6 Hz) and H‐24 (δH 3.53, dd, 12.4, 2.0 Hz), assigned by COSY and HMBC correlations. A third oxygenated methine resonance at δH 3.62 (ddd 12.6, 9.2, 4.0 Hz) was assigned to H‐7, based on correlations with H2‐6 (δH 1.68 m, 6α and 1.20 m, 6β) and H‐8 (δH 1.76, d 9.2 Hz) in the COSY spectrum, and with C‐5 (δC 46.8), C‐8 (δC 53.6), and C‐14 (δC 45.1) in the HMBC experiment. An HMBC cross‐peak between H‐16 and C‐24 defined the seven‐membered ring system with an ether bridge between C‐16 and C‐24, as in compounds 1 and 2. Additionally, a tertiary hydroxyl group at C‐25 (δC 73.0) was identified in compound 9 from the HMBC spectrum. NOESY correlations of H‐7 with H‐5, H‐6α, H‐15α, and CH3‐30 highlighted the β‐orientation of the hydroxy group at C‐7. The relative stereochemistry was confirmed by X‐ray analysis (Figure 2), and the absolute configuration was established based on its ECD spectrum (negative Cotton effect at λ 299 nm), similar to that of argentatin B (1), whose absolute configuration has been previously described [31].
FIGURE 2.
ORTEP drawing of compound 9.
2.2. In Silico Pesticide‐Likeness Prediction
The pesticide‐likeness properties of compounds 1–16 and betulinic acid (positive control), known for its activity against S. frugiperda [16], were calculated (Table 3). The calculated physicochemical parameters included molecular weight (MW < 500 uma), octanol/water coefficient (cLogP ‐1–3), number of hydrogen bond acceptors (HBAs 1–8), number of hydrogen bond donors (HBDs ≤ 2), number of rotatable bonds (RBs ≤ 12), and number of aromatic atoms (AA ≤ 17). These properties were analyzed based on the criteria outlined for identifying potential pesticide candidates [32].
TABLE 3.
Pesticide‐likeness prediction of compounds 1–16.
Compound | MW | cLogP | HBA | HBD | RB | AA | Tice violations |
---|---|---|---|---|---|---|---|
150–500 | 0–5 | 1‐8 | ≤2 | ≤12 | ≤17 | ≤1 | |
1 | 456.708 | 5.6514 | 3 | 1 | 1 | 0 | 1 |
2 | 458.724 | 5.5077 | 3 | 2 | 1 | 0 | 1 |
3 | 484.718 | 5.3454 | 4 | 0 | 2 | 0 | 1 |
4 | 470.691 | 4.9175 | 4 | 1 | 1 | 0 | 0 |
5 | 470.691 | 4.9175 | 4 | 1 | 1 | 0 | 0 |
6 | 398.585 | 4.6053 | 3 | 0 | 0 | 0 | 0 |
7 | 474.723 | 5.3259 | 4 | 3 | 5 | 0 | 2 |
8 | 490.722 | 4.4738 | 5 | 4 | 5 | 0 | 1 |
9 | 472.707 | 4.7993 | 4 | 2 | 1 | 0 | 0 |
10 | 302.456 | 4.1175 | 2 | 1 | 1 | 0 | 0 |
11 | 400.557 | 5.013 | 4 | 1 | 4 | 0 | 0 |
12 | 300.44 | 4.013 | 2 | 1 | 1 | 0 | 0 |
13 | 316.439 | 3.1609 | 3 | 2 | 1 | 0 | 0 |
14 | 314.423 | 3.3046 | 3 | 1 | 1 | 0 | 0 |
15 | 346.421 | 1.283 | 5 | 2 | 2 | 0 | 0 |
16 | 350.409 | 2.2523 | 6 | 1 | 3 | 0 | 0 |
Betulinic acid (Control) | 456.71 | 6.37 | 3 | 2 | 2 | 0 | 1 |
MW: Molecular weight; clogP: Octanol/water coefficient; HBA: Hydrogen bond acceptors; HBD: Hydrogen bond donors; RB: Rotatable bonds; AA: Aromatic atoms.
According to the Tice rules, a pesticide candidate should have an MW of less than 500 amu. In this study, the MW of the natural products isolated from V. dentata ranged from 484.72 to 300.44 amu. Compounds 1 and 2 had an MW similar to that of betulinic acid (control), while compounds 6, 10, and 12–16 had MWs lower than the control.
cLogP, which reflects the permeability of substances through cell membranes, was also assessed. Compounds 4–6, 8–10, and 12–16 had cLogP values of <5. Notably, compound 15 exhibited a lower cLogP (1.28) than the control (6.37). For insecticides, the mean value of cLogP typically ranges from 0.5 to 3.0 [32]. All the substances followed the HBA requirements proposed by Tice's rules, while compounds 1–6 and 9–16 fitted within the HBD interval.
In terms of the number of RB, Compounds 1, 2, 4–6, 9, 10, and 12–14 exhibited lower RB values than betulinic acid. A lower RB value indicates poor molecular flexibility, which is favorable for promoting major interactions with the target. Based on these analyses, 15 of the 16 (1–6 and 8–16) compounds were considered possible pesticide candidates, as they exhibited one or no violations of Tice's rules.
2.3. Evaluation of the Activity of the Isolated Compounds on the Inhibition of AChE
The inhibitory effects of the isolated compounds on AChE were also evaluated. Inhibiting this protein has been an effective strategy for insecticide development because it regulates the acetylcholine level, and specific residues in the insect AChE active sites are promising targets for the development of new insecticides [17]. Compounds showing a strong affinity for AChE can cause rapid paralysis and death in insects, making them candidates for pest control. The results showed weak activity for compounds 1 and 6, with IC50 values of 78.85 ± 6.95 and 53.54 ± 2.53 µM, respectively (Table 4). Previous studies on ent‐kaurenoic acid (10) reported IC50 values from 34.82 ± 0.45 µM [19] (using mouse brain AChE). In our conditions (using E. electricus AChE), compound 10 inhibited AChE by only 18.86% at 100 µg/mL during the primary screening (no IC50 was calculated). These discrepancies should be due to the different origins of AChE.
TABLE 4.
Inhibitory activity on acetylcholinesterase (AChE) of compounds 1–16, molecular docking analysis, and activity on Spodoptera frugiperda.
Activity on AchE | Molecular docking analysis | Activity on S. frugiperda | |||||
---|---|---|---|---|---|---|---|
Extract (compound/ligand) | 100 µM (%) | IC50 µM[ d ] | Binding energy (kcal/mol) | Dissociation constant [pM] | Larval weight in mg, (inhibition %)[ e ] | Mortality (%) | LC50 ppm[ f ] |
HE | 12.19[ a ] | nt | nc | nc | 72.4 ± 10 (70.74) | 14 ± 2.5 | 7397.85 |
AE | 13.80[ a ] | nt | nc | nc | 127.8 ± 12.9 (48.36) | 37 ± 9.8 | 2407.64 |
ME | 6.75 | nt | nc | nc | 106.4 ± 13.4 (56.99) | 32 ± 6.3 | 5362.7 |
1 | 67.18 | 78.85 ± 6.95 | −8.70 | 420 450.0 | 57.7 ± 7.3 (76.68) | 59 ± 10.3 | 51.61 |
2 | 15.60 | nt | −8.76 | 376 180.0 | 60.5 ± 7.2 (75.55) | 40 ± 9.1 | 84.56 |
3 | 20.95 | nt | −8.20 | 972 070.0 | nt | nt | nt |
4[ b ] | 10.87 | nt | −8.53 | 765 450.0 | nt | nt | nt |
5[ b ] | −7.96 | 1 470 000.0 | nt | nt | nt | ||
6 | 78.85 | 53.54 ± 2.53 | −8.89 | 306 530.0 | 84.1 ± 10.1 (66.02) | 17 ± 3.5 | 1433.4 |
7 | 36.78 | nt | −9.40 | 128 370.0 | nt | nt | nt |
8 | 4.25 | nt | −9.12 | 206 760.0 | nt | nt | nt |
9 | −7.03 | nt | −8.63 | 475 140.0 | 93.3 ± 10.3 (62.30) | 42 ± 8.7 | 99.66 |
10 | 18.86 | nt | −10.59 | 17 330.0 | 77.8 ± 7.3 (68.56) | 60 ± 12.4 | 24.69 |
11 | 26.05 | nt | −10.49 | 20 470.0 | nt | nt | nt |
12 | 26.41 | nt | −11.01 | 8550.0 | nt | nt | nt |
13 | 15.37 | nt | −11.59 | 3200.0 | 66.3 ± 6.5 (73.21) | 59 ± 17.8 | 62.04 |
14 | 2.35 | nt | −10.87 | 10 850.0 | nt | nt | nt |
15 | 9.11 | nt | −10.94 | 9560.0 | nt | nt | nt |
16 | 14.16 | nt | −11.14 | 6780.0 | nt | nt | nt |
Galantamine | 52.53[ c ] | 0.32 ± 0.01 | −10.32 | 27 300.0 | |||
Betulinic acid (control) | 71.55 [18] | 7.07 ± 1.09 [18] | −11.13 [18] | 2 450 000.0 | 129.22 ± 7.2 (47.78) | 47 ± 11.5 | 94.25 |
Artificial diet | 247.5 ± 22.4 | 7 ± 0.7 |
HE: hexane extract, AE: acetone extract, nt: not tested, nc: not calculated,
100 mg/L,
AChE activity tested as a mixture of 4 and 5,
galantamine reference compound 0.25 µM,
Half‐inhibitory concentration (IC50),
percentage of inhibition of larval weight in relation to the artificial diet control, and
Lethal concentration that killed 50% of the larvae (LC50).
2.4. Interactions between the Isolated Compounds and AChE Were Assessed Using Molecular Docking Techniques
Molecular docking calculations were performed to evaluate the interaction of ligands 1–16 with AChE. Residues involved in hydrogen bonding and hydrophobic interactions were identified, and the binding energy and dissociation constant were determined (Table 4).
ent‐Kaurenes 10 and 13 were found to be the most potent AChE inhibitors, with binding energies of −10.59 kcal/mol and −11.59 kcal/mol, respectively, and these relatively high energies correlated with the observed high mortality of S. frugiperda (50% lethal concentration for larval mortality [LC50] for 10: 24.69 ppm and LC50 for 13: 62.04 ppm; positive control: betulinic acid LC50 94.25 ppm) (See Table 4 and discussion below). Docking experiments for the ent‐kaurenoid acids 11, 12, 14, and 15 exhibited similar binding energies and dissociation constants, indicating strong affinity and specificity for the enzyme, but unfortunately, these compounds were not available for in vivo experiments.
The cycloartanes 1, 2, and 9 also exhibited relatively high binding energies of −8.70, −8.76, and −8.63 kcal/mol, respectively, in the docking experiments, and these values also correlated with the observed in vivo activity (LC50 for 1: 51.61 ppm, LC50 for 2: 84.56 ppm, and LC50 for 9: 99.66 ppm).
Figure 3 shows several significant molecular interactions for Argentatin B (1), ent‐kaurenoic acid (10) and 12α‐hydroxy ent‐kaur‐9(11),16‐dien‐19‐oic acid (13). Compound 1 interacts with AChE through hydrogen bonding interactions with P24 and Y101, these residues are not involved in the catalytic site, therefore an allosteric inhibition can explain the activity against the enzyme (Figure 3A). On the other hand, substances 10 and 13 interact with Y337 through hydrophobic contacts (Figure 3B,C; see Supporting Information). Y337 is responsible for maintaining the electrostatic balance of AChE's catalytic cavity, and this interaction is comparable to that observed with betulinic acid (the positive reference used in this work), which has been associated with AChE inhibition [16].
FIGURE 3.
Molecular docking interactions of some compounds tested against acetylcholinesterase (AChE). (A) Argentatin B (1), (B) ent‐kaurenoic acid (10), and (C) 12α‐hydroxy ent‐kaur‐9(11),16‐dien‐19‐oic acid (13).
Several amino acids participate in the contacts with the cycloartanes and the ent‐kaurenoic acids, including L285, W282, Q287, and E288 among others (see Supporting Information). The interactions shown in Figures 3A–C highlight the extensive range of molecular contacts that significantly influence the binding affinity of the compounds to AChE.
2.5. Insectistatic and Insecticide Activities of Extracts, and Compounds 1, 2, 6, 9, 10, and 13 on Spodoptera frugiperda
2.5.1. Insectistatic Effect
The insecticidal activity of the extracts and compounds 1, 2, 6, 9, 10, and 13 (selected based on their availability of 10 mg each for the assay) was tested against S. frugiperda. All extracts and compounds inhibited larval weight. Hexane, methanol, and acetone extracts reduced larval weight by 70.74%, 56.99%, and 48.36%, respectively, compared with the negative control (247.52 mg larval weight, Table 4). Compounds 1, 2, 6, 9, 10, and 13 clearly showed antifeedant properties showing higher percentages of inhibition of the larval weight in comparison with the reference betulinic acid (47.78% of inhibition) (Table 4).
2.5.2. Insecticide Effect
The results indicated that application of the acetone extract caused 37% mortality, followed by methanol and hexane extracts with 32% and 14% mortality, respectively. Among the pure compounds, ent‐kaurenoic acid (10) was the most active, followed by argentatin B (1), 12α‐hydroxy ent‐kaur‐9 (11),16‐dien‐19‐oic acid (13), argentatin D (2), and 7β‐hydroxy‐argentatin B (9). Their LC50 values were 24.69, 51.61, 62.04, 84.56, and 99.6 ppm, respectively. Betulinic acid (reference compound) had an LC50 of 94.25 ppm (Table 4). These activities are similar to those reported previously [13, 33, 34, 35].
In the in vitro AChE assays, cycloartane 6 exhibited weak activity (IC50 of 53.54 µM), followed by cycloartane 1 (IC50 78.85 µM). The same trend was observed in the docking analysis, with compound 6 showing a slightly lower binding energy (– 8.89 kcal/mol) than compound 1 (–8.70 kcal/mol). However, ent‐kaurane derivatives (10–15) and sesquiterpene lactone 16 were the most competitive inhibitors of AChE, with binding energies between –11.59 and –10.59 kcal/mol. These values were lower than that of galantamine, the reference compound, with a calculated binding energy of –10.32 kcal/mol (Table 4).
Compounds 1, 2, 9, 10, and 13 displayed LC50 mortalities values lower than the LC50 of the positive reference (betulinic acid 94.25 ppm) and can be considered bioactive compounds. The molecular docking results on AChE of compounds 10 and 13 correlated with those obtained from the evaluation against S. furgiperda: compound 10, whose affinity energy is among the highest (‐10.59 kcal/mol) presented the highest mortality of the substances evaluated (LC50 24.69 ppm), and compound 13 exhibited a binding energy of ‐11.59 kcal/mol and LC50 of 62.04 ppm. Compounds 1 (‐8.70 kcal/mol), 2 (‐8.76 kcal/mol), and 9 (‐8.63 kcal/mol) displayed LC50 of 51.61, 84.56, and 99.6 ppm, respectively. Thus, considering the mortality LC values, the relative activity of the isolated compounds was determined to be 10> 1> 13> 2> 9.
3. Conclusions
The chemical study of V. dentata, a species appreciated as a melliferous plant, led to the isolation of 20 compounds, including cycloartanes and ent‐kauranes as the main and characteristic secondary metabolites, with three new compounds identified (7–9). The majority of the isolated compounds were evaluated through in silico studies (pesticide‐likeness properties, indicating the absence of Tice's rules violations; molecular docking, recognizing the interactions of bioactive compounds with Y337 in the catalytic site of AChE); in vitro assays (inhibition of the enzyme AChE by some compounds), and in vivo studies (insectistatic and insecticide activities against S. frugiperda). Among the compounds evaluated, argentatin B (1) displayed the greatest inhibitory effect on S. frugiperda larval weight. This compound, along with cycloartanes 2 and 9, and the ent‐kaurenoic acids 10 and 13 had clear effects on insect mortality. The results of the docking calculations with the AChE enzyme correlated with the mortality results against the fall armyworm, and demonstrated that argentatins (1, 2, and 9) and ent‐kaurenoic acids (10 and 13) have the potential to control S. frugiperda.
4. Experimental
4.1. General Experimental Procedures
Melting points were determined on a Fisher‐Johns apparatus and were uncorrected. Optical rotations were obtained on a Perkin‐Elmer 343 polarimeter. IR spectra were recorded on a Thermo Scientific Nicolet iS50 FT‐IR spectrometer. ECD was obtained on a Jasco J‐720 CD spectropolarimeter. 1D and 2D NMR spectra were obtained on a Bruker Avance (F) 300 MHZ, a Bruker Avance III 400 MHz, or a Bruker AVANCE III HD 700 MHz spectrometer with tetramethylsilane (TMS) as the internal standard. X‐ray diffraction analysis was performed on an Xcalibur Atlas Gemini diffractometer with a Mo X‐ray source. The DART‐MS was performed using a JEOL AccuTOF JMS‐T100LC DART. Vacuum column chromatography (VCC) was performed under vacuum on silica gel G 60 (Merck, Darmstadt, Germany). Flash column chromatography (FCC) was performed on silica gel 230‐400 mesh (Macherey‐Nagel, Germany). Analytical TLC was performed on Si gel 60 GF254 or RP‐18 W/UV254 (10–40) µm, Macherey‐Nagel, Germany) and preparative TLC on Si gel GF254 layer thickness 2.0 mm or RP‐18 W/UV254 layer thickness 1.0 mm, using 10 × 20 cm plates.
4.2. Plant Material
Aerial parts of V. dentata (Cav.) Spreng. were collected along the road Tequisquiapan‐Queretaro, 30 km after Tequisquiapan (Hwy 200), Queretaro State, México, in October 2018 and authenticated by Prof. José Luis Villaseñor. A voucher specimen (MEXU 1473183) was deposited at the National Herbarium (Instituto de Biología, UNAM, México).
4.3. Spodoptera frugiperda (J.E. Smith)
The larvae of S. frugiperda were reared under laboratory conditions at the Departamento de Desarrollo de Productos Bióticos del Instituto Politécnico Nacional (Yautepec, Morelos). The diet formula was 800 mL distilled water, 60 g diet (Product No. F0635; S.W. Corn Borer, Bio‐Serv, Frenchtown, NJ, USA), 20 g sterile corn cob, 100 g ground corn, 40 g brewer's yeast, 10 g vitamins (Lepidoptera fortification blend, Bio‐Serv, Flemington, NJ), 10 g agar, 1.7 g sorbic acid, 1.7 g methyl p‐hydroxybenzoate, and 0.6 g neomycin sulfate. Insects were maintained in a climate chamber set at 25 ± 2°C, 60 ± 5% RH, and 12:12 h L:D [36].
4.4. Extraction and Isolation
The dried and ground aerial parts of V. dentata (800 g) were successively extracted with hexane, acetone, and MeOH at room temperature to obtain the respective extracts. The hexane extract (44 g) was fractionated on silica gel G 60 using VCC and eluted with a hexane‐EtOAc gradient system to obtain argentatin B (1, colorless prisms [hexane‐acetone] melting point [mp] 175–176°C, 1.2 g) and argentatin D (2, colorless prisms [hexane acetone] mp 234–235°C, 70 mg), from fractions eluted with hexane‐EtOAc 17:3 and 4:1, respectively, and mixtures A‐I. Mixture A (8.0 g), obtained with hexane‐EtOAc (19:1), after purification via a silica gel 230–400 mesh FCC eluted with hexane‐EtOAc 19‐1 afforded 1.8 g of a mixture of ent‐kaurenoic and grandiflorenic acids (4:1), which by recrystallization in methanol gave ent‐kaurenoic acid (10, colorless prisms [MeOH] mp 176–177°C, [α]25 D = –102, c = 0.13, EtOH, 560 mg), 60 mg of its mother liquors were separated by preparative RP TLC (MeOH‐H2O 4:1 × 2) to obtain 37 mg of 10 and 10 mg of grandiflorenic acid (12, colorless prisms [MeOH] mp 156–158°C, [α]25 D = + 33, c = 0.10, EtOH). Mixture B (2.76 g) obtained using hexane‐EtOAc 19‐1 gave mixtures B1 and B2. Mixture B1 (150 mg) was submitted to two successive FCCs (eluted with hexane‐EtOAc‐19‐1 and hexane‐acetone 19‐1, respectively) to afford 10 mg of 15‐angeloylgrandifloric (11, colorless prisms (hexane ‐acetone) mp 198–200°C, [α]25 D = –68, c = 0.13, CHCl3). Mixture B2 (372 mg) was purified by FCC (benzene‐acetone 97:3) to obtain verbenol (colorless oil, 27 mg). Mixture C (1.9 g), obtained with hexane‐EtOAc (9:1), was purified by FCC (hexane‐EtOAc 9:1) to yield compound 1 (92 mg) and the 1:1 mixture of β‐sitosterol and stigmasterol (45 mg). Mixture D (2.5 g), obtained with hexane‐EtOAc (17:3), was purified through an FCC (hexane‐EtOAc 9:1) to obtain 1 (320 mg). Mixture E (1.08 g) obtained with hexane‐EtOAc (17:3) was subjected to FCC (hexane‐acetone 9:1) to afford compounds 4 and 5 as a mixture (120 mg) and fraction E1. Fraction E1 (317 mg) by FCC (hexane‐acetone 17:3) gave 4 and 5 mixtures (22 mg), and 2 (10 mg). Mixture F (1.82 g) obtained with hexane‐EtOAc (4:1) was purified by FCC (hexane‐acetone 9:1) to yield 2 (80 mg). Mixture G (630 mg) obtained with hexane‐EtOAc 7:3 was purified by FCC (hexane‐acetone 17:3) to yield compound 2 (64 mg) and mixture G1. Mixture G1 (525 mg) after two successive FCC (CH2Cl2‐acetone 95:5 and hexane acetone 4:1, respectively) gave compound 14 (15 mg, colorless prisms (MeOH) mp: 128–130°C, = + 35, c = 0.10, CHCl3, 15 mg). Mixture H (1.6 g) eluted with hexane‐EtOAc 7:3 gave compound 13 (colorless needles (MeOH) mp: 187‐190°C, = + 83, c = 0.13, CHCl3, 262 mg) and mixture H1. Mixture H1 (395 mg) was purified by FCC (CH2Cl2‐acetone 19:1) to obtain a solid, which, by crystallization with isopropyl ether, afforded diversifolin (16, amorphous powder, 28 mg). Mixture I (964 mg) was eluted with hexane‐EtOAc 3:2 and purified by FCC (CH2Cl2‐acetone 9:1) to produce mixtures I1 and I2. Mixture I1 (189 mg) by FCC (CH2Cl2‐acetone 9:1) followed by preparative TLC (CH2Cl2‐acetone 17:3) afforded compound 9 (25 mg). Mixture I2 (195 mg) was purified by preparative TLC (hexane‐acetone 7:3) afforded compound 7 (13 mg). The acetone extract (43 g) was fractionated in a VCC (hexane‐EtOAc gradient system) to obtain mixtures J‐P. Mixture J (275 mg) obtained with hexane‐EtOAc 19:1 was purified with FCC (hexane‐acetone 19:1) to obtain caryophyllene oxide (colorless oil, 27 mg). Mixture K (920 mg), obtained with hexane‐EtOAc 9:1, was treated with charcoal/acetone followed by FCC (hexane‐EtOAc 4:1) to give compounds 1 (103 mg), 2 (31 mg), and 3 (18 mg). Mixture L (1.1 g), eluted with hexane‐EtOAc (17:3), was treated with activated charcoal/acetone to obtain an amber oil, which was submitted to a VCC eluted with a gradient of hexane‐EtOAc 19‐1 to 9:1 to obtain compounds 1 (15 mg), 2 (40 mg), and 6 (32 mg). Mixture M (209 mg) was treated with charcoal/acetone followed by FCC (hexane‐EtOAc 4:1) to obtain 6 (15 mg). Mixture N (4.3 g) obtained with hexane‐EtOAc (17:3) after charcoal/acetone treatment was purified by VCC (hexane‐acetone gradient system) to yield compound 13 (115 mg) and mixtures N1‐N2. Mixture N1 (610 mg) was subjected to FCC (hexane‐EtOAc 7:3) to obtain compounds 13 (56.7 mg) and 6 (5.5 mg). Mixture N2 (355 mg) was purified by FCC (hexane‐acetone 7:3) to yield compounds 13 (10 mg) and 15 (amorphous powder, [α]25 D =–16.4, c = 0.11, CHCl3, 16 mg). Mixture O (2.45 g), obtained with hexane‐EtOAc 4:1, after two successive FCCs (hexane‐EtOAc 3:2 and CH2Cl2‐acetone 4:1, respectively), yielded compound 9 (26 mg). Mixture P (3.8 g), obtained with hexane‐EtOAc 1:1, was purified by VCC (hexane‐EtOAc), followed by two successive FCC (EtOAc‐MeOH 9:1 and CH2Cl2‐MeOH 9:1, respectively) to obtain compound 8 (11 mg). From the methanol extract (60 g), a mixture of ent‐kaurenoic (10) and grandiflorenic acids (12, 1.5 g) was characterized, and β‐sitosteryl β‐D‐glucopyranoside (50 mg) and sucrose (80 mg) were isolated.
4.5. Spectral Data for the New Compounds
Argentatin B (1): CD (CHCl3) Δε max: + 0.233209, ‐ 0.321232, ‐ 3.242297 (c 2.2×10−3 M).
24‐Epi‐argentatin C (7): colorless needles (hexane‐acetone) mp 145–146°C, [α]25 D ‐8.2 (c 0.11, CHCl3); IR (ATR) ν max 3402, 1701 cm−1; CD (CHCl3) Δε max: +0.321213,–0.111233, –1.963299 (c 2.1×10−3 M); 1H NMR data, see Table 1; 13C NMR data, see Table 2; DART+ m/z 475 [M + H]+ (15), 457 (50), 439 (100), 421 (30); HRDARTMS m/z 475.37938 [M + H]+ (C30H51O4 requires 475.37873).
7β‐Hydroxy‐24‐epi‐argentatin C (8): white amorphous powder, [α]25 D + 5.8 (c 0.12, MeOH); IR (ATR) ν max 3352, 1702 cm−1; CD (CHCl3) Δεl max: ‐0.191217, ‐0.260237, ‐1432300 (c 2.4×10−3 M); 1H NMR data, see Table 1; 13C NMR data, see Table 2; DART+ m/z 491 [M + H]+ (10), 473 (25), 455 (90), 437 (100), 419 (30); HRDARTMS m/z 491.37583 [M + H]+ (C30H51O5 requires 491.37365).
7β‐Hydroxyargentatin B (9): colorless prisms (hexane‐acetone), mp 175–176°C, [α]25 D–62.5 (c 0.16, CHCl3); IR (ATR) ν max 3433, 1701 cm−1; CD (CHCl3) Δε max: + 0.072209, –0.420233,–3.725299, (c 3.4×10−3 M); 1H NMR data, see Table 1; 13C NMR data, see Table 2; DART+ m/z 473 [M + H]+ (10), 455 (50), 437 (100), 419 (30); HRDARTMS m/z 473.36445 [M + H]+ (C30H49O4 requires 473.36308).
Crystal data for compound 9: C30H48O4, Mr 472.68, monoclinic, space group P21, a = 15.673(4) Å, α = 90o, b = 6.0278(11) Å, β = 117.56(3)o, c = 15.800(4) Å; γ = 90o, V = 1323.2(6) Å3, Z = 2, Dc = 1.186 Mg/m3, F(000) = 520; crystal dimensions/shape/color 0.3560 × 0.1942 × 0.1406 mm3/prism/colorless. Reflections collected 9441, independent reflections 5333 [R(int) = 0.0718]; final R indices [I > 2s(I)] R1 = 0.0762, wR2 = 0.1611; R indices (all data) R = 0.1145, wR2 = 0.1907. Absolute structure parameter 1.5 (10). The deposition number 2445540 for compound 9 contains the supplementary crystallographic data for this work. These data are provided free of charge by the joint Cambridge Crystallographic Data Center and Fachinformationszentrum Karlsruhe http://www.ccdc.cam.ac.uk/structures Access Structures service.
Argentatin C (17): colorless prisms (hexane‐acetone) mp 175–176°C, [α]25 D + 5.7 (c 0.10, CHCl3); CD (CHCl3) Δε max: +0.219213,–0.129251, –1.547299 (c 2.5×10−3 M); 1H NMR data, see Table 1; 13C NMR data, see Table 2.
4.6. In Silico Prediction of Pesticide‐likeness Properties
The simplified molecular‐input line‐entry system format (SMILES) was obtained for compounds 1–16 and for betulinic acid (control). The chemical structures of 1–16 were drawn using ChemDraw software, subsequently transformed to SMILES format, and saved in a .csv Excel document. Then, this was exported to Data Warrior v.5.2.1 software [37] to calculate the following physicochemical properties: MW, cLogP, HBA, HBD, RBs, and the number of AAs.
4.7. In Vitro AChE Assay
The AChE inhibitory activity of the isolated compounds was determined by Ellman's method [38], as previously reported,[39] using AChE isolated from Electrophorus electricus. Primary screening of compounds 1–16 was performed using 1, 10, and 100 µM concentrations of each compound; samples with less than 50% inhibition at 100 µM were considered non‐active. Galantamine and betulinic acid were used as positive controls. The reported IC50 values are the average of five independent experiments.
4.8. Docking Methodology
The ligands were initially optimized using Gaussian16 software [40], employing the density functional theory (DFT) with the hybrid density functional B3LYP [41]. Subsequently, the FASTA file of the AChE, ID:1C20, protein of Electrophorus electricus was retrieved from the Protein Data Bank (PDB) [42]. For molecular homology and docking calculations, the YASARA [43] and WHAT IF [44] software packages were used with the AutoDockLG [45] algorithm. The molecular docking was performed specifically in one domain of AChE because the two domains are identical. A total of 50 docking runs were performed to evaluate the reproducibility and reliability of the results. This methodology aimed to identify the optimal interaction between the ligand and AChE via different conformations of the ligand. Analysis of hydrogen bonding and hydrophobic interactions, as well as visualization of ligand‐protein interactions, were conducted using PyMOL software [46].
4.9. Biological Activity Against Spodoptera frugiperda (J.E. Smith)
The insecticidal properties of methanol, hexane, and acetone extracts and of compounds 1, 2, 6, 9, 10, and 13 were evaluated in an artificial diet in neonatal larvae following the procedure described in the literature.[47, 48] The extracts and pure compounds were solubilized in methanol:dimethyl sulfoxide 95:5 and evaluated in a range of 500–2500 and 12.5–100 mg/kg (ppm), respectively. The experiments were carried out in plastic containers with lids measuring 3.0 × 3.5 cm in height and diameter; each larva was the experimental unit and replicated thirty times in triplicate. The dependent variables were the decrease in larval weight gain (mg) and larval mortality (%).
4.10. Statistical Analysis
The data were presented as the mean ± standard deviation. The test for normality (Shapiro–Wilk–W) and homoscedasticity (Bartlett test) were performed for all measured variables. One‐way to two‐way analysis of variance was performed to identify potential differences among treatments using Statistix 8.0 (Analytical Software, Florida, USA) [49]. Probit analysis was used to calculate the LC50 values using the JMP statistical software package ver. 11 [50].
Author Contributions
Amira Arciniegas: isolation and identification of compounds, and writing of the original draft. Olivia Pérez‐Valera: pest‐likeness predictions, data curation, and writing the final version. Simón Hernández Ortega: X‐ray determination of structure; Antonio Nieto Camacho, acetylcholinesterase assays and data curation. Israel Valencia: docking studies and review. Joel Daniel Castañeda‐Espinoza: Spodoptera frugiperda assays. Rodolfo Figueroa Brito: writing, review, infrastructure and resources for S. frugiperda assays. José Luis Villaseñor: collection, identification and registry of the plant and review. Guillermo Delgado: research design, data analysis, writing, review, infrastructure and resources.
Conflicts of Interest
The authors declare no conflicts of interest.
Supporting information
Supporting Information data associated with this article (1H, 13C, and 2D NMR spectra of compounds 7–9, X‐ray data, and docking results) are available on the www under https://doi.org/10.1002/MS‐number.
Acknowledgments
We are indebted to Adriana Romo Pérez, Ángeles Peña González, María Isabel Chávez Uribe, Beatriz Quiroz García, Javier Pérez Flores, Carmen García González, and Rubén Gaviño Ramírez, from the Institute of Chemistry, UNAM, and to Marcos Flores Alamo from the Faculty of Chemistry, UNAM for their technical assistance. This study made use of UNAM´s NMR lab: LURMN at IQ‐UNAM, which is funded by CONAHCYT‐ Mexico (Project 0224747) and UNAM. Israel Valencia acknowledges the use of the Miztli supercomputer with the LANCAD‐UNAM‐DGTIC‐049 and LANCAD‐UNAM‐DGTIC‐413 projects.
Open access funding provided by UNAM.
Arciniegas A., Pérez‐Valera O., Hernández‐Ortega S., et al. “New Terpenoids from Viguiera dentata: In Silico Pesticide‐Likeness Properties, Acetylcholinesterase Inhibition, Molecular Docking, and Evaluation against Spodoptera frugiperda .” Chemistry & Biodiversity 22, no. 9 (2025): 22, e202500917. 10.1002/cbdv.202500917
Funding: This work was supported by the Universidad Nacional Autónoma de México (Dirección General de Asuntos de Personal Académico, PAPIIT IG200821).
Data Availability Statement
Data supporting the findings are available from the corresponding author upon request.
References
- 1. Blake S. F., “A Revision of the Genus Viguiera ,” Contributions from the Gray Herbarium of Harvard University 54 (1918): 1–205. [Google Scholar]
- 2. Schilling E. E. and Panero J. L., “A Revised Classification of Subtribe Helianthinae (Asteraceae: Heliantheae) II. Derived Lineages,” Botanical Journal of the Linnean Society 167 (2011): 311–331, 10.1111/j.1095-8339.2011.01172.x. [DOI] [Google Scholar]
- 3. Arciniegas A., Pérez‐Castorena A. L., Romo de Vivar A., et al., “Secondary Metabolites in Viguiera (Compositae, Heliantheae, Helianthinae) and Segregated Genera. A Review of Their Biological Activities With Chemotaxonomic Observations,” Botanical Sciences 101 (2023): 1–40, 10.17129/botsci.3072. [DOI] [Google Scholar]
- 4. Turner B. L., “Recension of Viguiera (sensu stricto) (Asteraceae: Heliantheae) of Mexico,” Phytologia 97 (2015): 16–24. ISSN: 030319430. [Google Scholar]
- 5. Gao F., Mabry T. J., Bohlmann F., and Jakupovic J., “Cycloartanone Derivatives from Viguiera dentata ,” Phytochemistry 25 (1986): 1489–1491, 10.1016/S0031-9422(00)81319-9. [DOI] [Google Scholar]
- 6. Gao F., Miski M., Gage D. A., and Mabry T. J., “Terpenoid Constituents of Viguiera dentata ,” Journal of Natural Products 48 (1985): 316–318, 10.1021/np50038a021. [DOI] [PubMed] [Google Scholar]
- 7. Bohlmann F., Zdero C., and Mahanta P., “Neue Diterpene aus Dimorphotheca‐ und Viguiera‐arten,” Phytochemistry 16 (1977): 1073–1075, 10.1016/S0031-9422(00)86737-0. [DOI] [Google Scholar]
- 8. Bohlmann F., Jakupovic J., Ahmed M., et al., “Germacranolides and Diterpenes From Viguiera Species,” Phytochemistry 20 (1981): 113–116, 10.1016/0031-9422(81)85228-4. [DOI] [Google Scholar]
- 9. Canales M., Hernández T., Rodríguez‐Monroy M. A., et al., “Antimicrobial Activity of the Extracts and Essential Oil of Viguiera dentata ,” Pharmaceutical Biology 46 (2008): 719–723, 10.1080/13880200802215727. [DOI] [Google Scholar]
- 10. Cuevas‐Glory L. A., Ortiz‐Vazquez E., Pino J. A., and Sauri‐Duch E., “Floral classification of Yucatan Peninsula Honeys by PCA & HS‐SPME/GC–MS of Volatile Compounds,” International Journal of Food Science and Technology 47 (2012): 1378–1383, 10.1111/j.1365-2621.2012.02983.x. [DOI] [Google Scholar]
- 11. Cuevas‐Glory L., Pino J. A., and Sauri‐Duch E., “Volatile Constituents of Tahonal Flower (Viguiera dentata Blake, var. heliantoides) From the Yucatán Peninsula, México,” Journal of Essential Oil Research 20 (2008): 432–434, 10.1080/10412905.2008.9700050. [DOI] [Google Scholar]
- 12. Gao W., Dong X., Wei T., and Xing W., “The Chemical Structure and Bioactivity of Cycloartane‐Type Compounds,” Current Organic Chemistry 23 (2019): 2848–2872, 10.2174/1385272823666191203113221. [DOI] [Google Scholar]
- 13. Céspedes C. L., Martinez‐Vázquez M., Calderón J. S., Salazar J. R., and Aranda E., “Insect Growth Regulatory Activity of Some Extracts and Compounds From Parthenium argentatum on Fall Armyworm Spodoptera frugiperda,” Zeitschrift für Naturforschung C 56 (2001): 95–105, 10.1515/znc-2001-1-216. [DOI] [PubMed] [Google Scholar]
- 14. Wang L., Li D., Wang C., Zhang Y., and Xu J., “Recent Progress in the Development of Natural ent‐Kaurane Diterpenoids With Anti‐Tumor Activity,” Mini‐Reviews in Medicinal Chemistry 11 (2011): 910–919, 10.2174/138955711796575416. [DOI] [PubMed] [Google Scholar]
- 15. Simirgiotis M. J., García M., Sosa M. E., Giordano O. S., and Tonn C. E., “An ent‐Kaurene Derivative From Aerial Parts of Baccharis rufescens ,” Journal of the Argentine Chemical Society 91 (2003): 109–116, http://hdl.handle.net/11336/156926. [Google Scholar]
- 16. Pérez‐Valera O., Torres‐Martínez R., Nieto‐Camacho A., Valencia I., Javier Espinosa‐García F., and Delgado G., “Larvicidal Activity Against Spodoptera frugiperda of some Constituents From Two Diospyros Species. In Silico Pesticide‐Likeness Properties, Acetylcholinesterase Activity and Molecular Docking,” Chemistry and Biodiversity 21 (2024): 1–11, 10.1002/cbdv.202301871. [DOI] [PubMed] [Google Scholar]
- 17. Thapa S., Lv M., and Xu H., “Acetylcholinesterase: A Primary Target for Drugs and Insecticides,” Mini‐Reviews in Medicinal Chemistry 17 (2017): 1665–1676, 10.2174/1389557517666170120153930. [DOI] [PubMed] [Google Scholar]
- 18. Rodrigues G. C. S., dos Santos M. M., Silva C. A. B., de Sousa N. F., Scotti M. T., and Scotti L., “In Silico Studies of Lamiaceae Diterpenes With Bioinsecticide Potential Against Aphis gossypii and Drosophila melanogaster ,” Molecules 26 (2021): 766, 10.3390/molecules26030766. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19. Khan M.d. A. R., Islam M.d. A., Biswas K., et al., “Compounds From the Petroleum Ether Extract of Wedelia chinensis With Cytotoxic, Anticholinesterase, Antioxidant, and Antimicrobial Activities,” Molecules 28 (2023): 793, 10.3390/molecules28020793. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20. Jung H. A., Lee E. J., Kim J. S., et al., “Cholinesterase and BACE1 Inhibitory Diterpenoids From Aralia cordata ,” Archives of Pharmacal Research 32 (2009): 1399–1408, 10.1007/s12272-009-2009-0. [DOI] [PubMed] [Google Scholar]
- 21. Rodríguez‐Hahn L., Romo de Vivar A., Ortega A., Aguilar M., and Romo J., “Determinación de las estructuras de las argentatinas A, B y C del guayule,” Revista Latinoamericana de Quimica 1 (1970): 24–38. [Google Scholar]
- 22. Komoroski R. A., Gregg E. C., Shockcor J. P., and Geckle J. M., “Identification of Guayule Triterpenes by Two‐Dimensional and Multipulse NMR Techniques,” Magnetic Resonance in Chemistry 245 (1986): 534–543, 10.1002/mrc.1260240611. [DOI] [Google Scholar]
- 23. Romo de Vivar A., Martínez‐Vázquez M., Matsubara C., Pérez‐Sánchez G., and Joseph‐Nathan P., “Triterpenes in Parthenium argentatum, Structures of Argentatins C and D,” Phytochemistry 29 (1990): 915–918, 10.1016/0031-9422(90)80045-I. [DOI] [Google Scholar]
- 24. Enriquez R. G., Barajas J., Ortiz B., et al., “Comparison of Crystal and Solution Structures and 1 H and 13 C Chemical Shifts for Grandiflorenic Acid, Kaurenoic Acid, and Monoginoic Acid,” Canadian Journal of Chemistry 75 (1997): 342–347, 10.1139/v97-039. [DOI] [Google Scholar]
- 25. Ohno N., Mabry T. J., Zabelt V., and Watson W. H., “Tetrachyrin, a New Rearranged Kaurenoid Lactone, and Diterpene Acids from Tetrachyron orizabaensis and Helianthus debilis ,” Phytochemistry 18 (1979): 1687–1689, 10.1016/0031-9422(79)80184-3. [DOI] [Google Scholar]
- 26. Silva E. A., Takahashi J. A., Boaventura M. A. D., and Oliveira A. B., “The Biotransformation of Ent‐kaur‐16‐en‐19‐oic Acid by Rhizopus stolonifer ,” Phytochemistry 52 (1999): 397–400, 10.1016/S0031-9422(99)00219-8. [DOI] [PubMed] [Google Scholar]
- 27. Gao F., Wang H., and Mabry T. J., “Diterpenoids and a Sesquiterpene Lactone from Viguiera ladibractate ,” Phytochemistry 26 (1987): 779–781, 10.1016/S0031-9422(00)84785-8. [DOI] [Google Scholar]
- 28. Herz W., Kulanthaivel P., and Watanabe K., “ Ent‐kauranes and Other Constituents of Three Helianthus Species,” Phytochemistry 22 (1983): 2021–2025, 10.1016/0031-9422(83)80036-3. [DOI] [Google Scholar]
- 29. Bedir E., Çalis I., Aquino R., Piacente S., and Pizza C., “Cycloartane Triterpene Glycosides From the Roots of Astragalus brachypterus and Astragalus microcephalus ,” Journal of Natural Products 61 (1998): 1469–1472, 10.1021/np9801763. [DOI] [PubMed] [Google Scholar]
- 30. Ekiz G., Duman S., and Bedir E., “Biotransformation of Cyclocanthogenol by the Endophytic Fungus Alternaria eureka 1E1BL1,” Phytochemistry 151 (2018): 91–98, 10.1016/j.phytochem.2018.04.006. [DOI] [PubMed] [Google Scholar]
- 31. Parra‐Delgado H., Compadre C. M., Ramírez‐Apan T., et al., “Synthesis and Comparative Molecular Field Analysis (CoMFA) of Argentatin B Derivatives as Growth Inhibitors of Human Cancer Cell Lines,” Bioorganic & Medicinal Chemistry 14 (2006): 1889–1901, 10.1016/j.bmc.2005.10.038. [DOI] [PubMed] [Google Scholar]
- 32. Tice C. M., “Selecting the Right Compounds for Screening: Does Lipinski's Rule of 5 for Pharmaceuticals Apply to Agrochemicals?” Pest Management Science 57, no. 1 (2001): 3–16, 10.1002/1526-4998(200101)57:13.0.CO;2-6. [DOI] [PubMed] [Google Scholar]
- 33. Gutiérrez C., Gonzalez‐Coloma A., and Hoffmann J. J., “Antifeedant Properties of Natural Products From Parthenium argentatum, P. argentatum×P. tomentosum (Asteraceae) and Castela emoryi (Simaroubeaceae) against Reticulitermes flavipes ,” Industrial Crops and Products 10 (1999): 35–40, 10.1016/S0926-6690(99)00003-5. [DOI] [Google Scholar]
- 34. Jung H. A., Lee E. J., Kim J. S., et al., “Cholinesterase and BACE1 Inhibitory Diterpenoids From Aralia cordata ,” Archives of Pharmacal Research 32 (2009): 1399–1408, 10.1007/s12272-009-2009-0. [DOI] [PubMed] [Google Scholar]
- 35. Khan M.d. A. R., Islam M.d. A., Biswas K., et al., “Compounds From the Petroleum Ether Extract of Wedelia chinensis With Cytotoxic, Anticholinesterase, Antioxidant, and Antimicrobial Activities,” Molecules 28 (2023): 793, 10.3390/molecules28020793. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36. Burton R. L. and Perkins W. D., “WSB, A New Laboratory Diet for the Corn Earworm and the Fall Armyworm,” Journal of Economic Entomology 65 (1972): 385–386, 10.1093/jee/65.2.385. [DOI] [Google Scholar]
- 37. Sander T., Freyss J., Von Korff M., and Rufener C., “DataWarrior: An Open‐Source Program for Chemistry Aware Data Visualization and Analysis,” Journal of Chemical Information and Modeling 55 (2015): 460–473, 10.1021/ci500588j. [DOI] [PubMed] [Google Scholar]
- 38. Ellman G. L., Courtney K. D., Andres V., and Featherstone R. M., “A New and Rapid Colorimetric Determination of Acetylcholinesterase Activity,” Biochemical Pharmacology 7 (1961): 88–95. ISSN: 0006‐2952. [DOI] [PubMed] [Google Scholar]
- 39. Torres F. R., Pérez‐Castorena A. L., Arredondo L., et al., “Labdanes, Withanolides, and Other Constituents From Physalis nicandroides ,” Journal of Natural Products 82 (2019): 2489–2500, 10.1021/acs.jnatprod.9b00233. [DOI] [PubMed] [Google Scholar]
- 40. Frisch M. J., Trucks G. W., and Schlegel H. B., Gaussian 16, Revision B.01 (Gaussian, Inc., 2016). [Google Scholar]
- 41. Becke A. D., “Density‐Functional Thermochemistry. III. The Role of Exact Exchange,” Journal of Chemical Physics 98 (1993): 5648–5652, 10.1063/1.464913. [DOI] [Google Scholar]
- 42. Berman H. M., “The Protein Data Bank,” Nucleic Acids Research 28 (2000): 235–242, 10.1093/nar/28.1.235. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43. Land H. and Humble M. S., “YASARA: A Tool to Obtain Structural Guidance in Biocatalytic Investigations,” in Protein Engineering Methods and Protocols, eds. Bronscheuer U. T. and Höhner M. (Humana Press, 2018), 43–67, 10.1007/978-1-4939-7366-8_4. [DOI] [PubMed] [Google Scholar]
- 44. Vriend G., “WHAT IF: A Molecular Modeling and Drug Design Program,” Journal of Molecular Graphics and Modelling 8 (1990): 52–56, 10.1016/0263-7855(90)80070-v. [DOI] [PubMed] [Google Scholar]
- 45. Morris G. M., Huey R., Lindstrom W., et al., “AutoDock4 and AutoDockTools4: Automated Docking With Selective Receptor Flexibility,” Journal of Computational Chemistry 30 (2009): 2785–2791, 10.1002/jcc.21256. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46. DeLano W. L., The PyMOL Molecular Graphics System (De Lano Scientific, 2002). [Google Scholar]
- 47. Ramos‐López M. A., Pérez G. S., Rodríguez‐Hernández C., Guevara‐Fefer P., and Zavala‐Sánchez M. A., “Activity of Ricinus communis (Euphorbiaceae) Against Spodoptera frugiperda (Lepidoptera: Noctuidae),” African Journal of Biotechnology 9 (2010): 1359–1365, 10.5897/AJB10.1621. [DOI] [Google Scholar]
- 48. Figueroa‐Brito R., Tabares‐Parra A. S., Avilés‐Montes D., et al., “Chemical Composition of Jatropha curcas Seed Extracts and Its Bioactivity Against Copitarsia decolora under Laboratory and Greenhouse Conditions,” Southwestern Entomologist 46 (2021): 103–113, 10.3958/059.046.0110. [DOI] [Google Scholar]
- 49. Statistix 8: Analytical Software User's Manual (ScienceOpen, Inc., 2003). ISBN 1881789063. [Google Scholar]
- 50. SAS Institute, Inc . JMP 10. Basic Analysis and Graphing (JMP Statistical Discovery LLC, 2012), https://www.jmp.com/en_us/home.html. [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Supporting Information data associated with this article (1H, 13C, and 2D NMR spectra of compounds 7–9, X‐ray data, and docking results) are available on the www under https://doi.org/10.1002/MS‐number.
Data Availability Statement
Data supporting the findings are available from the corresponding author upon request.