Next Article in Journal
The Therapeutic Effect of Acanthopanax senticosus Components on Radiation-Induced Brain Injury Based on the Pharmacokinetics and Neurotransmitters
Previous Article in Journal
Gamma-Irradiation-Induced Degradation of the Water-Soluble Polysaccharide from Auricularia polytricha and Its Anti-Hypercholesterolemic Activity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and Insecticidal Activity of Fire Ant Venom Alkaloid-Based 2-Methyl-6-alkyl-Δ1,6-piperideines

1
State Key Laboratory of Integrated Management of Pest Insects and Rodents, Institute of Zoology, Chinese Academy of Sciences, Beijing 100101, China
2
School of Life Sciences, Institute of Life Science and Green Development, Hebei University, Baoding 071002, China
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(3), 1107; https://doi.org/10.3390/molecules27031107
Submission received: 23 December 2021 / Revised: 29 January 2022 / Accepted: 3 February 2022 / Published: 7 February 2022

Abstract

:
2,6-dialkylpiperideines found in the venom of Solenopsis (Hymenoptera, Formicidae) fire ants are a range of compounds possessing various biological activities. A series of racemic 2-methyl-6-alkyl-Δ1,6-piperideines were synthesized for chemical confirmation of the natural products found in fire ant venom, and the evaluation of their biological activity. Synthetic Δ1,6-piperideines and the natural compounds in the cis-alkaloid fraction of Solenopsis invicta had identical mass spectra and retention times. Their insecticidal activities against the third-instar larvae of cotton bollworm (Helicoverpa armigera) were evaluated by using injection and topical application methods. All three compounds exhibited no lethal effect at concentrations of 0.05–0.4 mol/L by topical treatment, but moderate lethal effect at 0.4 mol/L through injection treatment. Compound 6a showed significantly higher activity than the natural insecticide nicotine. The differences in activity among compounds 6b, 6c and nicotine were not significant. The elongation of the carbon chain at the 6-position of the piperideine ring appears to decrease insecticidal activity.

1. Introduction

The red imported fire ant, Solenopsis invicta, known as one of the most notorious invasive species in the world [1], has caused serious negative impacts on public health, agriculture and ecosystems [2,3,4,5,6]. It often becomes the dominant ant species in infested areas outside of its native range, due to its aggressive foraging behavior and high reproductive capability [7]. Having powerful venom may account for its aggressiveness and competitive advantage. For the purpose of understanding its taxonomic and toxicological properties, the chemistry of fire ant venom has been well studied [8,9]. Characteristic alkaloids are the major components of fire ant venom, while proteins make up only a small proportion [10]. The alkaloid component represented by the solenopsins contains predominately 2-methyl-6-alkyl or 2-methyl-6-alkenyl piperidines [10,11,12,13], with a series of minor piperideine alkaloids and pyridine alkaloids [14,15,16,17]. Those piperidine alkaloids have been demonstrated to possess diverse biological activities, including necrotic, hemolytic, antimicrobial and insecticidal activities [18,19,20,21]. Additionally, piperidine alkaloids can induce several different physiological activities in mammals [22,23,24,25,26,27] and serve as semiochemicals in insect communication [28,29,30].
The piperideine alkaloid from fire ant venom was first discovered in Solenopsis xyloni [10]. A mixture of 2-methyl-6-n-undecyl-Δ1,2-piperideine and 2-methyl-6-n-undecyl-Δ1,6-piperideine, obtained by the dehydration of cis-2-methyl-6-n-undecyl-piperidine with t-butylhypochlorite, has been used to establish the chemical identity of an odd peak found in Solenopsis xyloni. Reduction of this mixture with deuterated NaBH4 results in the disappearance of both peaks, and the presence of both cis- (major) and trans-2-methyl-6-n-undecyl-piperidine (minor). The reaction performed on the S. xyloni venom using NaBD4 gave the same results, which confirms the identity of 2-methyl-6-n-undecyl-Δ1,2-piperideine [10]. Subsequently, series of Δ1,2- and Δ1,6-piperideine alkaloids from the workers and alate queens of the red imported fire ants have been identified on the basis of their characteristic mass spectra [14,31,32,33]. However, to date, chemical confirmation of these piperideines with synthetic standards has not been reported.
2,6-dialkylpiperideines were proposed to function as precursors of the corresponding 2,6-dialkylpiperidines in their biosynthetic pathways [14,34,35]. Over the past decades, the antimicrobial and insecticidal properties of piperideine alkaloids have been reported by several groups [36,37,38,39,40,41]. 2,6-dialkylpiperideines have been demonstrated as a class of compounds that may exhibit interesting biological activities. However, piperideine alkaloid has remained a poorly studied topic compared to piperidine alkaloid. In the present study, we synthesized a series of racemic 2-methyl-6-alkyl-Δ1,6-piperideines and evaluated their insecticidal activity. The presence of 2-methyl-6-alkyl-Δ1,6-piperideines was confirmed with synthetic samples as well.

2. Results and Discussion

2.1. Chemistry

Racemic 2-methyl-6-alkyl-Δ1,6-piperideines were prepared following the synthetic route shown in Scheme 1 [39]. The large-scale preparation of glutarimide (2) was readily carried out using glutaric acid (1) and urea as starting materials, which reacted neat at 175 °C. One of the carbonyl groups on the piperidinone ring was converted to a methyl group in compound 3 by a Grignard reaction of glutarimide (2) with CH3MgBr reagent, and subsequent reduction with NaBH3CN. A key step was the generation of NBoc-aminoketones 5ac from Boc-protected compound 4, which was reacted with pre-prepared Grignard reagents to cause ring opening. Finally, the target compounds 6ac were obtained by catalytic cyclization of compounds 5ac with concentrated HCl. All copies of the spectra for compounds 24, 5ac, and 6ac are available in Supplementary Materials.

2.2. Confirmation of Δ1,6-Piperideines in Natural Venom

Fire ant venom alkaloids were readily extracted out of worker bodies and separated into cis- and trans-alkaloid fractions. The chemical identities of these alkaloids were determined by comparing the mass spectrum and the gas chromatographic retention time with published data. In both alkaloid fractions, two series of piperideines were isolated. Δ1,6-piperideines were identified by characteristic mass ions at m/z 96 and 111, whereas Δ1,2-piperideines were identified by characteristic mass ions at m/z 96, 97 and 110 [15,16]. Ultimately, synthetic compounds were directly compared with the natural alkaloids in the cis-alkaloid fraction, which have a well-established structure and stereochemistry. The retention times of Δ1,6-piperideine alkaloids from the natural venom are shown in Figure 1A. The mass spectra and retention times of synthetic compounds 6ac were completely consistent with those of the corresponding natural Δ1,6-piperideines. All three compounds had similar fragmentation patterns in the mass spectrum, with the intense ions at m/z 96 and 111 (Figure 1B–D). These data support the identification of Δ1,6-piperideine alkaloids in the venom of Solenopsis fire ants. The mass spectra of compounds 6ac matched those of the corresponding 2-methyl-6-undecyl-Δ1,6-piperideine, 2-methyl-6-tridecyl-Δ1,6-piperideine and 2-methyl-6-pentadecyl-Δ1,6-piperideine in the literature [10,14,15]. All these mass spectra showed the same base peak ion at m/z 111. However, Chen et al. [14] reported a slightly different mass pattern, with a base peak ion at m/z 96 for 2-methyl-6-tridecyl-Δ1,6-piperideine.

2.3. Toxicity against Cotton Bollworm (Helicoverpa armigera)

The insecticidal activities of compounds 6ac were evaluated by topical and injection treatments. As shown in Table 1, the mortality of piperideine 6b and 6c at various concentrations (0.05–0.4 mol/L) was similar to the standard control (nicotine) and the solvent control (acetone), indicating that these alkaloids were inactive against the early third-instar larvae of cotton bollworm at these test concentrations. Compound 6a, with a shorter side chain, exhibited a weak effect within 48 h, through topical application at concentrations of 0.2 and 0.4 mol/L.
For the injection treatment, a concentration of 0.4 mol/L was selected. The three piperideines exhibited moderate activities against the third-instar larvae of cotton bollworm (Table 2). Compound 6a, with an undecyl side chain, displayed the highest mortality rates of 45.5% (24 h) and 75.6% (48 h). Compounds 6b and 6c, with a longer side chain, showed relatively lower activity. The order of mortality from high to low is as follows: 6a > 6b > 6c, which means that the piperideines with a shorter side chain have better activity. Interestingly, as shown in Table 2, 6a displayed significantly higher activity in 48 h than nicotine, which has long been regarded as a natural insecticide [42]. At the same concentration, the toxicities of the piperideines in the subcutaneous injection were evidently greater than those in the topical administration, suggesting that the piperideines may possess weak penetrating capacity through the cuticle of the cotton bollworm larva. The significant difference in toxicity between topical and injection applications may explain why fire ant workers usually inject venomous poison into their prey while hunting.
The alkaloids in fire ant venom are mainly hydrophobic piperidines, along with piperideines in much lesser amounts. Piperideines have long been thought to be unstable intermediates for the biosynthesis of piperidines [10,15,35], but these remain largely unstudied. The facile syntheses of these Δ1,6-piperideines make it possible to study the biosynthesis of piperidines in fire ant venom. The mixed piperideine alkaloids from the extracts of the red imported fire ants show significant inhibitory effects on the growth of Pythium ultimum [36] and Clavibacter michiganensis subsp. michiganensis [41], and relatively high toxicity to the green peach aphid, Myzus persicae [38]. As they are minor components in fire ant venom, the availability of natural Δ1,6-piperideines is very limited. The synthetic version of Δ1,6-piperideines has the potential to control plant disease and pest insects. Future studies should focus on the scale-up synthesis of Δ1,6-piperideines and the development of Δ1,6-piperideine-based pesticides.

3. Materials and Methods

3.1. Synthesis

All reactions were carried out in dry glassware with a magnetic stirrer. Liquid solutions were transferred via syringe. All reactions were monitored by thin-layer chromatography on silica gel GF254 pre-coated plates (0.25 mm). The Grignard reagents n-C11H23MgBr, n-C13H27MgBr and n-C15H31MgBr were prepared according to a reported method from the corresponding n-alkyl bromide [43]. 1H and 13C NMR spectra were recorded on Varian Inova—400 MHz spectrometers using tetramethylsilane as the internal standard, and were reported in terms of chemical shift (δ ppm). The high-resolution mass spectra data were obtained using a Bruker mass spectrometer (Daltonik, Germany). Anhydrous THF was prepared by distillation over sodium with a diphenyl ketone indicator. Anhydrous DCM was collected by heating over CaH2. All other commercially obtained reagents and solvents were used as received.
Glutarimide (2): A mixture of glutaric acid (53.15 g, 0.40 mol) and urea (36.24 g, 0.60 mol) was melted for 40 min at 140 °C, and then the reaction mixture was stirred for an additional 2.5 h at 175 °C. Upon cooling, the crude product was recrystallized twice from EtOH to produce glutarimide 2 as a white solid (26.22 g, yield 58%) [44]. 1H NMR (400 MHz, CDCl3): δ 1.95–2.02 (m, 2H, CH2), 2.57 (t, 4H), 8.49 (s, 1H, NH).
6-Methyl-2-piperidinone (3): Into a stirred solution of glutarimide (15.00 g, 0.13 mol) in CH2Cl2 (100 mL) at −78 °C under a nitrogen atmosphere, the Grignard reagent CH3MgBr was added into 2-methyltetrahydrofuran (3 M, 130 mL) dropwise. Then the reaction mixture was stirred at room temperature for 12 h. NaBH3CN (9.83 g, 0.16 mol) was added, followed by the slow addition of a 6 N HCl solution to keep the pH at 3–4. The reaction was continued at 25 °C for 5 h. After completion, the mixture was neutralized with 1 N NaOH and extracted with CH2Cl2 (5 × 100 mL). The combined organic extracts were dried over anhydrous Na2SO4 and concentrated in vacuo. The crude product was purified by column chromatography on silica gel using ethyl acetate:MeOH (20:1~10:1) to obtain compound 3 as a white solid (9.75 g, yield 65%) [45]. 1H NMR (400 MHz, CDCl3): δ 1.20 (d, 3H, CH3), 1.30–1.37 (m, 1H), 1.66–1.74 (m, 1H), 1.87–1.91 (m, 2H), 2.24–2.31 (m, 1H), 2.35–2.41 (m, 1H), 3.48–3.54 (m, 1H), 5.96 (s, br, 1H).
6-Methyl-2-oxo-piperidine-1-carboxylic acid tert-butyl ester (4): To a solution of compound 3 (9.75 g, 0.086 mol), Et3N (13.07 g, 0.13 mol) and DMAP (0.52 g, 4.3 mmol) in CH2Cl2 (150 mL), (Boc)2O (28.12 g, 0.13 mol) was added. The solution was stirred to reflux for 18 h. When TLC showed that the reaction was complete, the solvent was evaporated and the resulting oil was purified via column chromatography on silica gel using hexanes; ethyl acetate (10:1) was used to obtain compound 4 as a yellow oil (12.86 g, yield 70%) [46]. 1H NMR (400 MHz, CDCl3): δ 1.27 (d, 3H, CH3), 1.52 (s, 9H), 1.66–1.77 (m, 2H), 1.90–1.96 (m, 2H), 2.43–2.51 (m, 2H), 4.26–4.31 (m, 1H).

3.1.1. General Procedure for Preparation of N-Boc-aminoketones (5ac):

To a stirred solution of tetramethylethylenediamine (TMEDA) (1.5 equiv.) and N-Boc-piperidinone (4) (1 equiv.) in anhydrous THF at −30 °C under a nitrogen atmosphere, the freshly prepared Grignard reagent RMgBr (1.5 equiv.) was added dropwise. The resulting suspension was allowed to be stirred at 25 °C for 10 h. After completion, 1N HCl was added to quench the reaction. The mixture was extracted with CH2Cl2. The combined organic phase was dried over anhydrous Na2SO4. The solvent was evaporated and the crude products were isolated by column chromatography on silica gel using hexanes; ethyl acetate (15:1) was used to afford the corresponding compounds.
Tert-butyl 6-oxoheptadecyl-2-carbamate (5a): Yield 45% from n-C11H23MgBr, white solid. 1H NMR (400 MHz, CDCl3): δ 0.88 (t, 3H, CH3), 1.12 (d, 3H, CH3), 1.21–1.31 (m, 16H, CH2), 1.37–1.44 (m, 2H, CH2), 1.44 (s, 9H, CH3 × 3), 1.54–1.61 (m, 4H, CH2), 2.36–2.44 (m, 4H, CH2), 3.59–3.66 (m, 1H, CH), 4.33 (s, br, 1H, NH); 13C NMR (125 MHz, CDCl3): δ 14.4 (CH3), 20.1 (CH3), 21.2 (CH2), 22.7 (CH2), 23.9 (CH2), 28.4 (CH3 × 3), 29.25 (CH2), 29.33 (CH2), 29.41 (CH2), 29.46 (CH2), 29.59 (CH2), 31.9 (CH2), 36.6 (CH2), 42.2 (CH2), 42.9 (CH2), 46.1 (CH), 79.0 (C), 155.4 (NHCO), 211.2 (CO).
Tert-butyl 6-oxononadecyl -2-carbamate (5b): Yield 41% from n-C13H27MgBr, white solid. 1H NMR (400 MHz, CDCl3): δ 0.89 (t, 3H, CH3), 1.12 (d, 3H, CH3), 1.21–1.32 (m, 22H, CH2), 1.44 (s, 9H, CH3 × 3), 1.53–1.66 (m, 4H, CH2), 2.34–2.55 (m, 4H, CH2), 3.59–3.68 (m, 1H, CH), 4.34 (s, br, 1H, NH); 13C NMR (125 MHz, CDCl3): δ 14.1 (CH3), 20.1 (CH3), 21.2 (CH2), 22.7 (CH2), 23.9 (CH2), 28.4 (CH3 × 3), 29.26 (CH2), 29.34 (CH2), 29.41 (CH2), 29.47 (CH2), 29.60 (CH2), 29.63 (CH2), 29.66 (CH2), 31.9 (CH2), 36.6 (CH2), 42.2 (CH2), 42.9 (CH2), 46.1 (CH), 79.0 (C), 155.4 (NHCO), 211.2 (CO).
Tert-butyl 6-oxohenicosyl -2-carbamate (5c): Yield 32% from n-C15H27MgBr, white solid [39]. 1H NMR (400 MHz, CDCl3): δ 0.80 (t, 3H, CH3), 1.03 (d, 3H, CH3), 1.12–1.26 (m, 26H, CH2), 1.35 (s, 9H, CH3 × 3), 1.44–1.57 (m, 4H, CH2), 2.28–2.35 (m, 4H, CH2), 3.52–3.56 (m, 1H, CH), 4.44 (s, br, 1H, NH); 13C NMR (125 MHz, CDCl3): δ 14.0 (CH3), 20.0 (CH3), 21.1 (CH2), 22.6 (CH2), 23.4 (CH2), 28.3 (CH3 × 3), 29.18 (CH2), 29.29 (CH2), 29.35 (CH2), 29.41 (CH2), 29.55 (CH2), 29.59 (CH2), 29.62 (CH2), 31.9 (CH2), 36.5 (CH2), 42.1 (CH2), 42.8 (CH2), 46.0 (CH), 78.8 (C), 155.4 (NHCO), 211.2 (CO).

3.1.2. General Procedure for Preparation of 2-Methyl-6-alkyl-Δ1,6-piperideines (6ac)

To a stirred solution of N-Boc-aminoketone (5ac) (0.5 g) in CH2Cl2 (10 mL) at 0 °C, 12 M HCl (5 mL) was added dropwise gradually over 15 min. After stirring for 5 h at 0 °C, the reaction mixture was basified with 2 M NaOH and the resulting mixture was extracted with CH2Cl2. The combined organic phase was dried over anhydrous Na2SO4. After workup, the crude product was purified by column chromatography on silica gel using hexanes; ethyl acetate (3:1) was used to produce the 2-methyl-6-alkyl-Δ1,6-piperideines.
2-Methyl-6-undecyl-Δ1,6-piperideine (6a): Yield 85% from 5a [47]. 1H NMR (400 MHz, CDCl3): δ 0.63 (t, 3H, CH3), 0.99 (d, 3H, CH3), 1.01–1.05 (m, 16H, CH2), 1.27–1.36 (m, 4H, CH2), 1.46–1.54 (m, 2H, CH2), 1.79–1.98 (m, 4H, CH2), 3.21 (m, 1H, CH); 13C NMR (125 MHz, CDCl3): δ 13.8 (CH3), 18.2 (CH3), 22.5 (CH2), 22.9 (CH2), 26.8 (CH2), 28.2 (CH2), 29.01 (CH2), 29.13 (CH2), 29.25 (CH2), 29.33 (CH2), 29.40 (CH2), 31.7 (CH2), 40.7 (CH2), 52.6 (CH), 171.3 (C=N); HRMS calculated for C17H33N [M + H]+ = 252.2691, found 252.2685.
2-Methyl-6-tridecyl-Δ1,6-piperideine (6b): Yield 77% from 5b. 1H NMR (400 MHz, CDCl3): δ 0.67 (t, 3H, CH3), 0.90–1.06 (m, 23H, CH2, CH3), 1.24–1.38 (m, 4H, CH2), 1.51–1.54 (m, 2H, CH2), 1.83–1.96 (m, 4H, CH2), 3.21 (m, 1H, CH); 13C NMR (125 MHz, CDCl3): δ 13.9 (CH3), 18.4 (CH3), 22.5 (CH2), 23.1 (CH2), 26.8 (CH2), 28.2 (CH2), 29.16 (CH2), 29.19 (CH2), 29.31 (CH2), 29.38 (CH2), 29.49 (CH2), 29.52 (CH2), 31.7 (CH2), 41.0 (CH2), 52.7 (CH), 170.1 (C=N); HRMS calculated for C19H37N [M + H]+ = 280.3004, found 280.2997.
2-Methyl-6-pentadecyl-Δ1,6-piperideine (6c): Yield 52% from 5c [39]. 1H NMR (400 MHz, CDCl3): δ 0.83 (t, 3H, CH3), 1.10–1.23 (m, 27H, CH2, CH3) 1.46–1.55 (m, 4H, CH2), 1.67–1.74 (m, 2H, CH2), 2.04–2.19 (m, 4H, CH2), 3.43 (m, 1H, CH); 13C NMR (125 MHz, CDCl3): δ 14.1 (CH3), 18.3 (CH3), 22.6 (CH2), 23.0 (CH2), 27.0 (CH2), 28.4 (CH2), 29.09 (CH2), 29.32 (CH2), 29.42 (CH2), 29.49 (CH2), 29.58 (CH2), 29.61 (CH2), 29.65 (CH2), 31.9 (CH2), 40.8 (CH2), 52.7 (CH), 171.4 (C=N); HRMS calculated for C21H41N [M + H]+ = 308.3317, found 308.3311.

3.2. Comparison of Retention Times and Mass Spectra of Synthetic Δ1,6-Piperideines with Those of Corresponding Natural Products

3.2.1. Source of Ant Colony

Workers of the red imported fire ant, S. invicta, were collected from Guangzhou, China (113.51° E, 23.18° N). The colony was maintained under laboratory conditions (24–26 °C, relative humidity 60% and 14L:10D h photoperiod) in a 1-gallon plastic jar coated with a Teflon suspension to prevent escape, and was fed sugar water and crickets. Ant workers were subjected to extraction within one month of collection.

3.2.2. Extraction, Isolation and Identification of Δ1,6-Piperideines

The extraction and isolation of venom alkaloids from S. invicta workers were conducted following previously reported procedures [15,16,48]. Briefly, the hexane extract of ant workers (1 g) randomly selected from the colony, was concentrated to 0.5 mL under a mild stream of nitrogen and loaded onto a 13 mm i.d. gravity chromatography column (20 g of silica gel, 300–400 mesh). The sample was sequentially eluted with hexane and a series of mixtures of hexane and acetone containing 2% triethylamine. Each 2 mL of eluent was collected, and its chemistry was analyzed by a Shimadzu GC–2030 equipped with a flame ionization detector (FID). The dimensions of the capillary column used were as follows: HP–5MS, 30 m × 0.25 mm i.d., 0.25 μm (Agilent). The injector was operated at 250 °C in splitless mode with the split opened after 1 min. Nitrogen was used as the carrier (1 mL/min) and makeup gas. The GC program used was as follows: injection at 90 °C (isothermal for 2 min), increase at 15 °C/min up to 270 °C and hold for 16 min. The detector temperature was set at 280 °C. In terms of appearance of characteristic peaks and their retention times in the GC chromatograms of each collection, the cis- and trans-alkaloid fractions were obtained by pooling collections containing specific groups of alkaloids. A mixture of synthetic Δ1,6-piperideines 6a, 6b and 6c in equal amounts was dissolved in hexane. To confirm the identities of Δ1,6-piperideines, both the cis-alkaloid fraction with a final volume of 1.5 mL and the solution of synthetic standards were subjected to GC-MS analysis using a Shimadzu GCMS-TQ8050 NX equipped with a Shimadzu SH-Rxi-5Sil MS column with exactly the same dimensions as the HP-5MS column above. The oven temperature was programmed as follows: 90 °C for 1 min, increased at 10 °C/min to 160 °C, then to 250 °C at 3 °C/min, and held for 2 min. The transfer line temperature was set at 250 °C. Mass spectra were obtained using electron impact (EI, 70 eV). Helium was used as the carrier gas (1 mL/min).

3.3. Insecticidal Activity Bioassays

3.3.1. Insect Rearing Procedure

Eggs of Helicoverpa armigera were purchased from a commercial supplier, Keyun Industry Co., Ltd. (Jiyuan, China). Newly hatched H. armigera larvae were reared in groups of 50–100 in 200 mL plastic cups at 27 ± 1 °C, relative humidity 60% and 16L:8D h photoperiod, until the third instar. Small cubes of artificial diet [49] were refreshed daily. The early third-instar larvae were individually transferred to 9 cm-diameter Petri dishes and used for bioassays. The in vivo effect of piperideines 6ac was determined through topical and injection applications.

3.3.2. Topical Application

Piperideines 6ac and nicotine were dissolved in acetone and then diluted to obtain the following concentrations: 0.4, 0.2, 0.1 and 0.05 mol/L. Acetone was used as a negative control, whereas nicotine was used as a positive control. A 2 μL drop of acetone solution was applied with a micropipette onto the thoracic dorsum of an anesthetized early third-instar larva placed in a metal bath at 4 °C. Twenty larvae were treated for each treatment. Treated larvae were placed individually in a 10 mL glass tube and maintained on an artificial diet during the post-treatment observation period under the standard rearing conditions describe above. Larvae were considered dead if they were unable to move in a coordinated manner when touched several times with a hairbrush. The numbers of dead larvae were recorded 24 h and 48 h after treatment for the calculation of mortality.

3.3.3. Injection Application

Injections of 6ac and nicotine solutions at 0.4 mol/L were performed using the Nanoject II Injector (Drummond Scientific Company, Broomall, PA, USA). The needle of the injector was carefully inserted between the second and third abdominal segments of an anesthetized third-instar larva placed in the icy metal bath at a vertical angle of about 30° under a dissecting microscope. A total of 202.4 nL (50.6 nL × 4) of the sample was applied to a single larva. Sixteen larvae were treated in each treatment. The tests were replicated five times over 5 consecutive days. Acetone was served as the control. The treated larvae were maintained separately under standard conditions. Mortality was recorded 24 h and 48 h after treatment.

4. Conclusions

In summary, three minor components of the fire ant venom and 2-methyl-6-alkyl-Δ1,6-piperideines were synthesized and used to confirm the identification of the corresponding natural products. Their retention times and mass spectra data coincided with those in the cis-alkaloid fraction purified from the hexane extract of S. invicta workers. Their insecticidal activities were further evaluated with topical application and injection methods. The results of the bioassays indicated that these compounds displayed insecticidal properties against the early third-instar larvae of H. armigera. In particular, compound 6a, with an undecyl side chain, displayed promising injection activity as compared to the natural alkaloid nicotine at 0.4 mol/L. The activity of 2-methyl-6-alkyl-Δ1,6-piperideines is negatively correlated to the length of the carbon chain at the 6-position of the piperideine ring. Interestingly, the piperideines were found to be almost inactive when applied topically. The cuticle of the cotton bollworm may prevent the penetration of piperideines into the haemolymph, resulting in such inactivity by topical application. The insecticidal activities of compounds 6ac provide valuable information for further research on biological activities.

Supplementary Materials

The following supporting information can be downloaded online: Figures S1–S3: 1H NMR of 24, Figures S4–S9: 1H NMR and 13C NMR of 5ac, Figures S10–S18: 1H NMR, 13C NMR, and HRMS of 6ac.

Author Contributions

X.W. and L.C. conceived and designed the experiments, analyzed the data, and wrote the paper. X.W., G.W. and G.X. performed the experiments. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China (grant No. 31772540) and the High-level Talents Research Start-up Project of Hebei University (grant No. 521000981387).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are available from the authors.

References

  1. Lowe, S.; Browne, M.; Boudjelas, S.; De Poorter, M. 100 of the World’s Worst Invasive Alien Species—A Selection from the Global Invasive Species Database; The Invasive Species Specialist Group: Auckland, New Zealand, 2000. [Google Scholar]
  2. Vinson, S.B. Impact of the invasion of the imported fire ant. Insect. Sci. 2013, 20, 439–455. [Google Scholar] [CrossRef] [PubMed]
  3. Ascunce, M.S.; Yang, C.-C.; Oakey, J.; Calcaterra, L.; Wu, W.-J.; Shih, C.-J.; Goudet, J.; Ross, K.G.; Shoemaker, D. Global invasion history of the fire ant Solenopsis invicta. Science 2011, 331, 1066–1068. [Google Scholar] [CrossRef]
  4. Adams, C. Agricultural and medical impact of the imported fire ants. In Fire Ants and Leaf-Cutting Ants: Biology and Management; Lofgren, C.S., Vander Meer, R.K., Eds.; CRC Press: Boca Raton, FL, USA, 1986; pp. 48–57. [Google Scholar]
  5. Wang, L.; Xu, Y.J.; Zeng, L.; Lu, Y.Y. Impact of the red imported fire ant Solenopsis invicta Buren on biodiversity in South China: A review. J. Integr. Agric. 2019, 18, 788–796. [Google Scholar] [CrossRef]
  6. Wojcik, D.P.; Allen, C.R.; Brennan, R.J.; Forys, E.A.; Jouvenaz, D.P.; Lutz, R.S. Red imported fire ants: Impact on biodiversity. Am. Ent. 2001, 47, 24–50. [Google Scholar] [CrossRef]
  7. Allen, C.R.; Epperson, D.M.; Garmestani, A.S. Red imported fire ant impacts on wildlife: A decade of research. Am. Midl. Nat. 2004, 152, 88–103. [Google Scholar] [CrossRef] [Green Version]
  8. MacConnell, J.G.; Blum, M.S.; Buren, W.F.; Williams, R.N.; Fales, H.M. Fire ant venoms: Chemotaxonomic correlations with alkaloidal compositions. Toxicon 1976, 14, 69–78. [Google Scholar] [CrossRef]
  9. Chen, J.; Shang, H. Advances in research on the venom chemistry of imported fire ants. In Recent Advances in Entomological Research: From Molecular Biology to Pest Management; Liu, T., Kang, L., Eds.; Springer: Berlin/Heidelberg, Germany, 2011; pp. 417–433. [Google Scholar]
  10. Brand, J.M.; Blum, M.S.; Fales, H.M.; MacConnell, J.G. Fire ant venoms: Comparative analyses of alkaloidal components. Toxicon 1972, 10, 259–271. [Google Scholar] [CrossRef]
  11. MacConnell, J.G.; Blum, M.S.; Fales, H.M. Alkaloid from fire ant venom: Identification and synthesis. Science 1970, 168, 840–841. [Google Scholar] [CrossRef]
  12. Leclercq, S.; Thirionet, I.; Broeders, F.; Daloze, D.; Vander Meer, R.K.; Braekman, J.C. Absolute configuration of the solenopsins, venom alkaloids of the fire ants. Tetrahedron 1994, 50, 8465–8478. [Google Scholar] [CrossRef]
  13. Jones, T.H.; Blum, M.S.; Fales, H.M. Ant venom alkaloids from Solenopsis and Monomorium species: Recent developments. Tetrahedron 1982, 38, 1949–1958. [Google Scholar] [CrossRef]
  14. Chen, J.; Cantrell, C.L.; Shang, H.-W.; Rojas, M.G. Piperideine alkaloids from the poison gland of the red imported fire ant (Hymenoptera: Formicidae). J. Agric. Food Chem. 2009, 57, 3128–3133. [Google Scholar] [CrossRef] [PubMed]
  15. Chen, L.; Fadamiro, H.Y. Re-investigation of venom chemistry in Solenopsis fire ants. I. Identification of novel alkaloids in S. richteri. Toxicon 2009, 53, 469–478. [Google Scholar] [CrossRef] [PubMed]
  16. Chen, L.; Fadamiro, H.Y. Re-investigation of venom chemistry in Solenopsis fire ants. II. Identification of novel alkaloids in S. invicta. Toxicon 2009, 53, 479–486. [Google Scholar] [CrossRef] [PubMed]
  17. Chen, J.; Zhao, Y.; Li, X.-C.; Zhao, J.-H. Pyridine alkaloids in the venom of imported fire ants. J. Agric. Food Chem. 2019, 67, 11388–11395. [Google Scholar] [CrossRef]
  18. Attygalle, A.B.; Morgan, E.D. Chemicals from the glands of ants. Chem. Soc. Rev. 1984, 13, 245–278. [Google Scholar] [CrossRef]
  19. Jouvenaz, D.P.; Blum, M.S.; MacConnell, J.G. Antibacterial activity of venom alkaloids from the imported fire ant, Solenopsis invicta Buren. Antimicrob. Agents Chemother. 1972, 2, 291–293. [Google Scholar] [CrossRef] [Green Version]
  20. Tadeusz, A. Alkaloids: Chemistry, Biology, Ecology, and Applications, 2nd ed.; Elsevier B.V.: Amsterdam, The Netherlands, 2015; p. 496. [Google Scholar]
  21. Escoubas, P.; Blum, M.S. The biological activities of ant-derived alkaloids. Appl. Myrmecol. World Perspect. 1990, 482–489. [Google Scholar]
  22. Yi, G.B.; McClendon, D.; Desaiah, D.; Goddard, J.; Lister, A.; Moffitt, J.; Vander Meer, R.K.; Deshazo, R.; Lee, K.S.; Rockhold, R.W. Fire ant venom alkaloid, isosolenopsin A, a potent and selective inhibitor of neuronal nitric oxide synthase. Int. J. Toxicol. 2003, 22, 81–86. [Google Scholar] [CrossRef]
  23. Javors, M.A.; Zhou, W.; Maas, J.W.; Han, S.; Keenan, R.W. Effects of fire ant venom alkaloids on platelet and neutrophil function. Life Sci. 1993, 53, 1105–1112. [Google Scholar] [CrossRef]
  24. Arbiser, J.L.; Kau, T.; Konar, M.; Narra, K.; Ramchandran, R.; Summers, S.A.; Vlahos, C.J.; Ye, K.Q.; Perry, B.N.; Matter, W.; et al. Solenopsin, the alkaloidal component of the fire ant (Solenopsis invicta), is a naturally occurring inhibitor of phosphatidylinositol-3-kinase signaling and angiogenesis. Blood 2007, 109, 560–565. [Google Scholar] [CrossRef] [Green Version]
  25. Yeh, J.Z.; Narahashi, T.; Almon, R.R. Characterization of neuromuscular blocking action of piperidine derivatives. J. Pharmacol. Exp. Ther. 1975, 194, 373. [Google Scholar]
  26. Read, G.W.; Lind, N.K.; Oda, C.S. Histamine release by fire ant (Solenopsis) venom. Toxicon 1978, 16, 361–367. [Google Scholar] [CrossRef]
  27. Foster, D.; Ahmed, K. Effect of piperidines and fire ant venom on ATPase activities from brain homogenate fractions and characterization of Na+-K+ ATPase inhibition. Biochem. Pharmacol. 1977, 26, 983–985. [Google Scholar] [CrossRef]
  28. Chen, L.; Sharma, K.R.; Fadamiro, H.Y. Fire ant venom alkaloids act as key attractants for the parasitic phorid fly, Pseudacteon tricuspis (Diptera: Phoridae). Naturwissenschaften 2009, 96, 1421–1429. [Google Scholar] [CrossRef]
  29. Eliyahu, D.; Ross, K.; Haight, K.; Keller, L.; Liebig, J. Venom alkaloid and cuticular hydrocarbon profiles are associated with social organization, queen fertility status, and queen genotype in the fire ant Solenopsis invicta. J. Chem. Ecol. 2011, 37, 1242–1254. [Google Scholar] [CrossRef] [Green Version]
  30. Sharma, K.R.; Fadamiro, H.Y. Fire ant alarm pheromone and venom alkaloids act in concert to attract parasitic phorid flies, Pseudacteon spp. J. Insect Physiol. 2013, 59, 1119–1124. [Google Scholar] [CrossRef]
  31. Chen, L.; Lu, Y.-Y.; Hu, Q.-B.; Fadamiro, H.Y. Similarity in venom alkaloid chemistry of alate queens of imported fire ants: Implication for hybridization between Solenopsis richteri and S. invicta in the Southern United States. Chem. Biodivers. 2012, 9, 702–713. [Google Scholar] [CrossRef]
  32. Chen, J.; Shang, H.; Jin, X. Interspecific variation of Δ1,6-piperideines in imported fire ants. Toxicon 2010, 55, 1181–1187. [Google Scholar] [CrossRef]
  33. Chen, L.; Hu, Q.-B.; Fadamiro, H.Y. Reduction of venom alkaloids in Solenopsis richteri × Solenopsis invicta hybrid: An attempt to identify new alkaloidal components. J. Agric. Food Chem. 2010, 58, 11534–11542. [Google Scholar] [CrossRef]
  34. Yu, Y.-T.; Wei, H.-Y.; Fadamiro, H.Y.; Chen, L. Quantitative analysis of alkaloidal constituents in imported fire ants by gas chromatography. J. Agric. Food Chem. 2014, 62, 5907–5915. [Google Scholar] [CrossRef]
  35. Leclercq, S.; Braekman, J.C.; Daloze, D.; Pasteels, J.M.; Vander Meer, R.K. Biosynthesis of the solenopsins, venom alkaloids of the fire ants. Naturwissenschaften 1996, 83, 222–225. [Google Scholar] [CrossRef]
  36. Li, S.; Jin, X.; Chen, J. Effects of piperidine and piperideine alkaloids from the venom of red imported fire ants, Solenopsis invicta Buren, on Pythium ultimum Trow growth in vitro and the application of piperideine alkaloids to control cucumber damping-off in the greenhouse. Pest Manag. Sci. 2012, 68, 1546–1552. [Google Scholar] [CrossRef] [PubMed]
  37. Lai, L.-C.; Chang, Y.-Y.; Hua, K.-H.; Wu, W.-J.; Huang, R.-N. Comparative toxicity of three fire ant (Hymenoptera: Formicidae) venoms to Spodoptera litura larvae. Sociobiology 2010, 56, 653–663. [Google Scholar]
  38. Rashid, T.; Chen, J.; McLeod, P. Toxicity of newly isolated piperideine alkaloids from the red imported fire ant, Solenopsis invicta Buren, against the green peach aphid, Myzus persicae (Sulzer). Adv. Entomol. 2013, 1, 20–23. [Google Scholar] [CrossRef]
  39. Yan, Y.; An, Y.; Wang, X.; Chen, Y.; Jacob, M.R.; Tekwani, B.L.; Dai, L.; Li, X.-C. Synthesis and antimicrobial evaluation of fire ant venom alkaloid based 2-methyl-6-alkyl-Δ1,6-piperideines. J. Nat. Prod. 2017, 80, 2795–2798. [Google Scholar] [CrossRef]
  40. Dai, L.; Jacob, M.R.; Khan, S.I.; Khan, I.A.; Clark, A.M.; Li, X.-C. Synthesis and antifungal activity of natural product-based 6-alkyl-2,3,4,5-tetrahydropyridines. J. Nat. Prod. 2011, 74, 2023–2026. [Google Scholar] [CrossRef] [Green Version]
  41. Li, S.; Jin, X.; Chen, J.; Lu, S. Inhibitory activities of venom alkaloids of red imported fire ant against Clavibacter michiganensis subsp. michiganensis in vitro and the application of piperidine alkaloids to manage symptom development of bacterial canker on tomato in the greenhouse. Int. J. Pest Manag. 2013, 59, 150–156. [Google Scholar]
  42. Tomizawa, M.; Casida, J.E. Neonicotinoid insecticide toxicology: Mechanisms of selective action. Annu. Rev. Pharmacol. Toxicol. 2004, 45, 247–268. [Google Scholar] [CrossRef] [Green Version]
  43. Maiti, M.; Roy, A.; Roy, S. Effect of pH and oxygen atom of the hydrophobic chain on the self-assembly property and morphology of the pyridyl boronic acid based amphiphiles. Langmuir 2013, 29, 13329–13338. [Google Scholar] [CrossRef]
  44. Annese, C.; D’Accolti, L.; Fusco, C.; Licini, G.; Zonta, C. Heterolytic (2e) vs homolytic (1e) oxidation reactivity: N−H versus C−H switch in the oxidation of lactams by dioxirans. Chem. Eur. J. 2017, 23, 259–262. [Google Scholar] [CrossRef]
  45. Li, H.; Wu, H.; Zhang, H.; Su, Y.; Yang, S.; Hensen, E.J.M. A facile direct route to N-(Un)substituted lactams by cycloamination of oxocarboxylic acids without external hydrogen. ChemSusChem 2019, 12, 3778–3784. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Maldaner, A.O.; Pilli, R.A. Stereoselective alkylation of N-Boc-2-pyrrolidinones and N-Boc-2-piperidinones. Synthesis and characterization of disubstituted lactams. Tetrahedron 1999, 55, 13321–13332. [Google Scholar] [CrossRef]
  47. Meyers, A.I.; Edwards, P.D.; Bailey, T.R.; Jagdmann, G.E., Jr. α-Amino carbanions. Preparation, metalation, and alkylation of enamidines. Synthesis of piperidine and pyrrolidine natural products and homologation of carbonyl compounds. J. Org. Chem. 1985, 50, 1019–1026. [Google Scholar] [CrossRef]
  48. Shi, Q.-H.; Hu, L.; Wang, W.-K.; Vander Meer, R.K.; Porter, S.D.; Chen, L. Workers and alate queens of Solenopsis geminata share qualitatively similar but quantitatively different venom alkaloid chemistry. Front. Ecol. Evol. 2015, 3, 76. [Google Scholar] [CrossRef] [Green Version]
  49. Mironidis, G.K.; Savopoulou-Soultani, M. Development, survivorship, and reproduction of Helicoverpa armigera (Lepidoptera: Noctuidae) under constant and alternating temperatures. Environ. Entomol. 2008, 37, 16–28. [Google Scholar] [CrossRef]
Scheme 1. Synthesis of 2-methyl-6-alkyl-Δ1,6-piperideines. Reagents and conditions: () urea, neat, 175 °C, 2.5 h; () CH3MgBr/CH2Cl2, −78~25 °C, 12 h; () NaBH3CN, pH 3~4, r.t, 5 h; () (Boc)2O, Et3N, DMAP, CH2Cl2, reflux, 18 h; () RMgBr, TMEDA, THF, −30~25 °C, 12 h; () 12M HCl, CH2Cl2, 0 °C, 5 h.
Scheme 1. Synthesis of 2-methyl-6-alkyl-Δ1,6-piperideines. Reagents and conditions: () urea, neat, 175 °C, 2.5 h; () CH3MgBr/CH2Cl2, −78~25 °C, 12 h; () NaBH3CN, pH 3~4, r.t, 5 h; () (Boc)2O, Et3N, DMAP, CH2Cl2, reflux, 18 h; () RMgBr, TMEDA, THF, −30~25 °C, 12 h; () 12M HCl, CH2Cl2, 0 °C, 5 h.
Molecules 27 01107 sch001
Figure 1. Chemical confirmation of Δ1,6-piperideines. (A) GC-MS chromatographs of the cis-alkaloid fraction and a mixture of compounds 6ac. (BD) Mass spectra of compounds 6ac.
Figure 1. Chemical confirmation of Δ1,6-piperideines. (A) GC-MS chromatographs of the cis-alkaloid fraction and a mixture of compounds 6ac. (BD) Mass spectra of compounds 6ac.
Molecules 27 01107 g001
Table 1. Contact toxicities of 6ac against the early third-instar larvae of cotton bollworm following topical application.
Table 1. Contact toxicities of 6ac against the early third-instar larvae of cotton bollworm following topical application.
CompoundConcentration (mol/L)Mortality (%)
24 h48 h
6a0.415.015.0
0.25.015.0
0.100
0.0500
6b0.45.05.0
0.200
0.100
0.0500
6c0.400
0.205
0.100
0.0500
Nicotine0.45.010.0
0.200
0.100
0.0500
Control (Acetone)-00
Table 2. Insecticidal activities of 6ac against the early third-instar larvae of cotton bollworm following injection application.
Table 2. Insecticidal activities of 6ac against the early third-instar larvae of cotton bollworm following injection application.
CompoundMortality (%) (Mean ± SD) A
24 h48 h
6a B45.5 ± 15.9 d75.6 ± 5.7 e
6b B27.6 ± 12.8 bc48.5 ± 24.9 d
6c B20.1 ± 10.1 b37.2 ± 2.4 cd
Nicotine B32.9 ± 10.2 bcd43.7 ± 6.0 d
Acetone1.3 ± 2.8 a2.5 ± 3.4 a
A The means followed by different letters indicate significant difference (p < 0.05) in same column by the Tukey-HSD test. B The concentration of test compounds and nicotine is 0.4 mol/L.
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, X.; Wang, G.; Xu, G.; Chen, L. Synthesis and Insecticidal Activity of Fire Ant Venom Alkaloid-Based 2-Methyl-6-alkyl-Δ1,6-piperideines. Molecules 2022, 27, 1107. https://doi.org/10.3390/molecules27031107

AMA Style

Wu X, Wang G, Xu G, Chen L. Synthesis and Insecticidal Activity of Fire Ant Venom Alkaloid-Based 2-Methyl-6-alkyl-Δ1,6-piperideines. Molecules. 2022; 27(3):1107. https://doi.org/10.3390/molecules27031107

Chicago/Turabian Style

Wu, Xiaoqing, Guangyu Wang, Guangxin Xu, and Li Chen. 2022. "Synthesis and Insecticidal Activity of Fire Ant Venom Alkaloid-Based 2-Methyl-6-alkyl-Δ1,6-piperideines" Molecules 27, no. 3: 1107. https://doi.org/10.3390/molecules27031107

Article Metrics

Back to TopTop