Next Article in Journal
Chemo-Enzymatic Synthesis of Silybin and 2,3-Dehydrosilybin Dimers
Previous Article in Journal
Simultaneous Determination of Original, Degraded Ginsenosides and Aglycones by Ultra High Performance Liquid Chromatography Coupled with Quadrupole Time-of-Flight Mass Spectrometry for Quantitative Evaluation of Du-Shen-Tang, the Decoction of Ginseng
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fluorescent Probes for Insect Ryanodine Receptors: Candidate Anthranilic Diamides

1
Department of Applied Chemistry, College of Science, China Agricultural University, No. 2 Yuanmingyuan West Road, Beijing 100193, China
2
College of Agriculture and Biotechnology, China Agricultural University, No. 2 Yuanmingyuan West Road, Beijing 100193, China
3
Department of Environmental Science, Policy, and Management, University of California, Berkeley, CA 94720-3112, USA
*
Author to whom correspondence should be addressed.
Molecules 2014, 19(4), 4105-4114; https://doi.org/10.3390/molecules19044105
Submission received: 16 January 2014 / Revised: 13 March 2014 / Accepted: 21 March 2014 / Published: 2 April 2014

Abstract

:
Diamide insecticides with high efficacy against pests and good environmental safety are broadly applied in crop protection. They act at a poorly-defined site in the very complex ryanodine (Ry) receptor (RyR) potentially accessible to a fluorescent probe. Two N-propynyl analogs of the major anthranilic diamide insecticides chlorantraniliprole (Chlo) and cyantraniliprole (Cyan) were accordingly synthesized and converted into two fluorescent ligands by click reaction coupling with 3-azido-7-hydroxy-2H-chromen-2-one. The new diamide analogs and fluorescent ligands were shown to be nearly as potent as Chlo and Cyan in inhibition of [3H]Chlo binding and stimulation of [3H]Ry binding in house fly thoracic muscle RyR. Although the newly synthesized compounds had only moderate activity in insect larvicidal activity assays, their high in vitro potency in a validated insect RyR binding assay encourages further development of fluorescent probes for insect RyRs.

1. Introduction

Natural ryanodine (Ry) from Ryania speciosa was used as a botanical insecticide in the 1940s [1]. By the late 1990s, two types of diamides were found to possess excellent insecticidal activity, producing poisoning signs similar to those of Ry. Radioligand binding and Ca2+ flux studies at Nihon Nohyaku and Bayer [2,3,4,5,6,7,8,9] and at DuPont [10,11,12,13] defined the mode of action of the diamides as Ry receptor (RyR) modulators keeping Ca2+ channels open and depleting Ca2+ stores, leading to gradual muscle contraction and paralysis. RyRs form intracellular Ca2+ channels located mostly on the sarcoplasmic reticulum of muscle. With commercialization of the anthranilic diamides chlorantraniliprole (Chlo) [11] and cyantraniliprole (Cyan) [14] (Figure 1) and phthalic diamide flubendiamide (Flu) [2], the insect RyR became one of the most important insecticide targets. Photoaffinity labeling revealed that Flu interacts in the insect transmembrane domain and that the N-terminus plays an important role in sensitivity [8]. Radioligand binding assays with [3H]Chlo and [3H]Ry established that anthranilic diamides and Ry act at the Musca domestica RyR [15,16]. The genes encoding the RyRs have been characterized for various species (Heliothis virescens, Myzus persicae, Aphis gossypii, Peregrinus maidis and Drosophila melanogaster) [10], and more recently for Plutella xylostella [17]. However, up to now, very little is known about the structure of the insect tetrameric transmembrane RyRs and particularly their active binding site.
Figure 1. Anthranilic diamide skeleton common to two major commercial insecticides and four analogs with N-propynyl and N-hydroxycoumarin substituents.
Figure 1. Anthranilic diamide skeleton common to two major commercial insecticides and four analogs with N-propynyl and N-hydroxycoumarin substituents.
Molecules 19 04105 g001
This study takes the first steps in developing a fluorescent probe for the insect RyR. Chlo and Cyan were selected as prototypes because they are exceptionally potent and readily recognizable insecticides. 7-Hydroxycoumarin was used as the fluorescent substituent to be introduced. Retention of the required biological properties was established by high potency in a validated insect (Musca) RyR radioligand binding assay and moderate larvicidal activity on an important pest (Spodoptera).

2. Results and Discussion

2.1. Fluorescent Ligands Design and Synthesis

The key step in preparing the effective fluorescent ligands was solved by coupling of Chlo and Cyan as the pharmacophore skeletons and hydroxycoumarin as the fluorescent chromophore with a click reaction (Scheme 1). Both fluorescent ligands obtained are very stable and possess good fluorescent properties, for instance, both their maximum excitation wavelength 396 nm and the maximum emission wavelength 475 nm and their large stokes shift are suitable for fluorescence assay.
Scheme 1. Synthesis routes for two N-propynyl intermediates and two fluorescent ligands.
Scheme 1. Synthesis routes for two N-propynyl intermediates and two fluorescent ligands.
Molecules 19 04105 g003

2.2. Fluorescent Ligands Bind Potently to Specific Site in Musca RyR

The validated Musca RyR assay with [3H]Chlo and [3H]Ry was used to compare the potencies of the new compounds with the standards Chlo and Cyan. The results showed that Fluor-Cl and Fluor-CN stimulated [3H]Ry binding with EC50’s of 5–7 nM (Figure 2A) consistent with the reported value of 3–15 nM for Chlo and Cyan [15]. In addition, Fluor-Cl and Fluor-CN inhibited [3H]Chlo binding with similar IC50’s of 34–39 nM (Figure 2B), compared to 6–14 nM for Chlo and Cyan [15], which indicated that they all work at the same binding site on Musca RyRs.
Figure 2. Effects of fluorescent ligands on specific binding of [3H]Ry and [3H]Chlo in Musca muscle membrane. (A) Stimulation of [3H]Ry binding (EC50). (B) Inhibition of [3H]Chlo binding (IC50).
Figure 2. Effects of fluorescent ligands on specific binding of [3H]Ry and [3H]Chlo in Musca muscle membrane. (A) Stimulation of [3H]Ry binding (EC50). (B) Inhibition of [3H]Chlo binding (IC50).
Molecules 19 04105 g002

2.3. Insecticidal Activity of Fluorescent Ligands

The results of assays with Plutella larvae are shown in Table 1. Although the two N-propynyl compounds were less potent than Chlo, Cyan and Flu, they all showed good insecticidal activity with LC50s for Prop-Cl and Prop-CN of 0.08 ppm and 0.34 ppm, respectively. After incorporating the hydroxycoumarin fluorophore, the ligands were 22–32 fold less potent than the propynyl compounds, with LC50s of 2.6–7.6 ppm. Thus the new compounds retain good insecticidal activity despite the propynyl and large polar fluorescent substituents were introduced. Most importantly the fluorescent probes are very potent in vitro in binding assays, although less toxic in vivo with penetration barriers.
Table 1. Insecticidal activity for Plutella larvae.
Table 1. Insecticidal activity for Plutella larvae.
CompdLC50 (95%CL) a (ppm)Slope SEx2 (df) b
Chlo0.033(0.006–0.059)2.5 ± 0.513.7(6)
Cyan0.061(0.035–0.086)5.0 ± 1.014.5(8)
Flu0.056(0.037–0.068)6.9 ± 2.11.7(7)
Prop-Cl0.080(0.024–0.152)1.5 ± 0.35.5 (7)
Prop-CN0.34(0.24–0.43)4.6 ± 1.17.9(8)
Fluor-Cl2.59(1.69–3.46)4.0 ± 0.95.2(8)
Fluor-CN7.57(4.57–11.77)2.4 ± 0.73.8(8)
a 95% confidence limits, n = 30; b Chi-square value and degrees of freedom (df) as calculated by POLOPlus.

3. Experimental

3.1. General Methods and Materials

Commercial reagents were used as obtained from Sigma-Aldrich (St. Louis, MO, USA), Alfa Aesar (Ward Hill, MA, USA), and J&K (Beijing, China). Technical grade Chlo (95% purity) was obtained from DuPont Agricultural Chemicals Ltd. (Shanghai, China). Sources for Cyan and other relevant chemicals were reported earlier [15,16]. Hepes was produced in-house. Nuclear magnetic resonance (NMR) spectra were recorded with a 300 MHz spectrometer (Bruker, Billerica, MA, USA) and mass spectra (MS) with a time-of-flight spectrometer (Agilent Technologies, Inc., Santa Clara, CA, USA). Fluorescence intensity was measured with a Cary Eclipse Fluorescence Spectrophotometer (Agilent). 6-Chloro-8-methyl-1H-benzo[d][1,3]oxazine-2,4-dione (2-Cl), 8-methyl-2,4-dioxo-2,4-dihydro-1H-benzo[d][1,3]oxazine-6-carbonitrile (2-CN) and 3-bromo-1-(3-chloropyridin-2-yl)-1H-pyrazole-5-carboxylic acid (5) were prepared from compounds 1 and 4, respectively, by the methods of Chai et al. [18] and 3-azido-7-hydroxy-2H-chromen-2-one (7) was synthesized according to Sivakumar et al. [19]. The structures of compounds are shown in Scheme 1.

3.2. Synthesis of Propynyl Intermediates

3-Bromo-1-(3-chloropyridin-2-yl)-1H-pyrazole-5-carbonyl chloride (6). Compound 5 (3.66 g, 12 mmol, 1 eq) was placed in a 100-mL round-bottomed flask, and SOCl2 (4.43 g, 37.5 mmol, 3.12 eq) was added dropwise under stirring at room temperature. After addition, the mixture was heated gently to 100 °C and kept for 4 h. The reaction mixture was cooled to room temperature and then excess SOCl2was distilled (70 °C) at atmospheric pressure to obtain 3.80 g of compound 6 as a deep colored oil with yield 98%.
2-Amino-5-chloro-3-methyl-N-(prop-2-ynyl)benzamide (3-Cl). Acetic acid (7.37 g, 44 mM, 3.67 eq) was added to a solution of compound 2-Cl (6.0 g, 30 mmol, 2.5 eq) and ethyl acetate (50 mL). Propynylamine (2 g, 36 mmol, 3 eq) was added, the reaction mixture was heated and kept at 50 °C for 1 h and then was cooled to room temperature. Water (65 g, 3.6 mol, 300 eq) was added and the solvent was removed by distillation under vacuum. The residue was cooled below 10 °C with an ice bath, forming a precipitate, which was filtered, washed with water (approx. 10 mL), and dried to afford 7.0 g of compound 3-Cl as a white powder in 96% yield.
3-Bromo-N-(4-chloro-2-methyl-6-(prop-2-ynylcarbamoyl)phenyl)-1-(3-chloropyridin-2-yl)-1H-pyrazole-5-carboxamide (Prop-Cl). A mixture of compound 6 (9.0 g, 28 mmol, 2.33 eq) and compound 3-Cl (6 g, 27 mmol, 2.25 eq) in 1,2-dichloroethane (90 mL) was cooled to 10 °C with an ice bath, and the reaction solution was stirred at room temperature for 2 h then N,N-diisopropylethylamine (4.2 g, 30 mmol, 2.5 eq) was added dropwise. When the reaction was completed, based on thin layer chromatography (TLC) monitoring, water (50 mL) was added and the reaction mixture was stirred for 15 min. The white solid precipitate was collected by filtration, washed with water (3 mL), and dried to afford 9.2 g of compound Prop-Cl as a white powder, yield 92%; m.p. 222.5–223.6 °C, 1H-NMR (DMSO-d6): δ 10.24 (s, 1H), 8.79 (t, J = 5.4 Hz, 1H), 8.49 (dd, J = 4.7, 1.5 Hz, 1H), 8.16 (dd, J = 8.1, 1.5 Hz, 1H), 7.60 (dd, J = 8.1, 4.7 Hz, 1H), 7.50 (d, J = 2.0 Hz, 1H), 7.38 (s, 1H), 7.33 (d, J = 2.3 Hz, 1H), 3.93 (dd, J = 5.4, 2.5 Hz, 2H), 3.09 (t, J = 2.5 Hz, 1H), 2.16 (s, 3H); 13C-NMR (DMSO-d6): δ 165.81, 156.05, 148.85, 147.51, 139.81, 139.67, 139.44, 135.70, 132.07, 131.97, 131.46, 128.31, 127.22, 127.00, 126.00, 111.21, 81.12, 73.55, 28.98, 18.11; HRMS (ESI) calcd for C20H14BrCl2N5O2 [M+H]+ 507.9786, found 507.9806.
2-Amino-5-cyano-3-methyl-N-(prop-2-ynyl)benzamide (3-CN) and 3-bromo-N-(4-cyano-2-methyl-6-(prop 2-ynylcarbamoyl)phenyl)-1-(3-chloropyridin-2-yl)-1H-pyrazole-5-carboxamide (Prop-CN). These compounds were synthesized according to the procedures for compounds 3-Cl and Prop-Cl respectively. Prop-CN: a white powder, yield 90%; m.p. 223.7–225.0 °C, 1H-NMR (DMSO-d6): δ 10.49 (s, 1H), 8.89 (t, J = 5.4 Hz, 1H), 8.50 (dd, J = 4.7, 1.5 Hz, 1H), 8.16 (dd, J = 8.1, 1.5 Hz, 1H), 7.90 (d, J = 1.3 Hz, 1H), 7.75 (d, J = 1.5 Hz, 1H), 7.61 (dd, J = 8.1, 4.7 Hz, 1H), 7.40 (s, 1H), 3.94 (dd, J = 5.3, 2.6 Hz, 2H), 3.11 (t, J = 2.5 Hz, 1H), 2.21 (s, 3H); 13C-NMR (DMSO-d6): δ 165.58, 155.88, 148.75, 147.56, 139.73, 139.54, 138.44, 137.73, 135.99, 134.58, 130.20, 128.29, 127.28, 127.07, 118.41, 111.45, 109.81, 80.94, 73.77, 29.08, 18.14; HRMS (ESI) calcd for C21H14BrClN6O2 [M+H]+ 499.0130, found 499.0144.

3.3. Synthesis of Fluorescent Ligands

3-Bromo-N-(4-chloro-2-((1-(7-hydroxy-2-oxo-2H-chromen-3-yl)-1H-1,2,3-triazol-4-yl)me-thylcarbamoyl)-6-methylphenyl)-1-(4-chloropyridin-2-yl)-1H-pyrazole-5-carboxamide (Fluor-Cl). To a solution of Prop-Cl (0.25 g, 0.5 mmol) and compound 7 (0.1 g, 0.5 mmol) in water and THF (10 mL, v/v = 1:1), freshly prepared aqueous sodium ascorbate (150 μL, 0.15 mmol) and aqueous copper (II) sulfate pentahydrate (125 μL, 0.038 mmol) were added in order. The heterogeneous mixture was stirred vigorously overnight in the dark at room temperature. The solvent THF was removed by distillation, the residue was diluted with water (5 mL), cooled in an ice bath, and then the precipitate was collected by filtration, washed with cold water (10 mL) followed by drying under vacuum to afford 0.3 g of pure Fluor-Cl as a yellow powder, yield 85.7%; m.p. 212.3–213.8 °C, 1H-NMR (DMSO-d6): δ 10.90 (s, 1H), 10.27 (s, 1H), 8.96 (t, J = 5.8 Hz, 1H), 8.53 (s, 1H), 8.44 (dd, J = 4.7, 1.5 Hz, 1H), 8.34 (d, J = 10.4 Hz, 1H), 8.13 (dd, J = 8.1, 1.5 Hz, 1H), 7.75 (d, J = 8.6 Hz, 1H), 7.57 (dd, J = 8.1, 4.7 Hz, 1H), 7.50 (d, J = 1.8 Hz, 1H), 7.39 (d, J = 2.3 Hz, 1H), 7.28 (s, 1H), 6.91 (dd, J = 8.5, 2.2 Hz, 1H), 6.86 (d, J = 2.1 Hz, 1H), 4.48 (d, J = 5.5 Hz, 2H), 2.16 (s, 3H); 13C-NMR (DMSO-d6): δ 166.22, 162.84, 156.69, 156.16, 155.10, 148.75, 147.41, 145.39, 139.74, 139.65, 139.38, 136.59, 136.24, 131.96, 131.83, 131.49, 131.40, 128.19, 127.16, 126.96, 125.94, 124.19, 119.77, 114.71, 111.02, 110.84, 102.60, 67.47, 18.07; HRMS (ESI) calcd for C29H20BrCl2N8O5 [M+H]+ 711.0119, found 711.0118.
3-Bromo-1-(3-chloropyridin-2-yl)-N-(4-cyano-2-((1-(7-hydroxy-2-oxo-2H-chromen-3-yl)-1H-1,2,3-triazol-4-yl)methylcarbamoyl)-6-methylphenyl)-1H-pyrazole-5-carboxamide (Fluor-CN). This compound was synthesized according to the procedure for Fluor-Cl. A yellow powder, yield 71.4%, m.p. 234.7–236.9 °C 1H-NMR (DMSO-d6): δ 10.90 (s, 1H), 10.53 (s, 1H), 9.05 (t, J = 5.6 Hz, 1H), 8.54 (s, 1H), 8.45 (dd, J = 4.7, 1.3 Hz, 1H), 8.39 (s, 1H), 8.13 (dd, J = 8.1, 1.3 Hz, 1H), 7.90 (s, 1H), 7.82 (d, J = 1.4 Hz, 1H), 7.75 (d, J = 8.6 Hz, 1H), 7.58 (dd, J = 8.1, 4.7 Hz, 1H), 7.32 (s, 1H), 6.92 (dd, J = 8.5, 2.2 Hz, 1H), 6.86 (d, J = 2.1 Hz, 1H), 4.50 (d, J = 5.5 Hz, 2H), 2.22 (s, 3H); 13C-NMR (DMSO-d6): δ 165.63, 162.55, 156.37, 155.68, 154.79, 148.34, 147.14, 144.93, 139.39, 139.19, 138.09, 137.32, 136.35, 135.54, 134.83, 131.08, 129.88, 127.86, 126.91, 126.70, 123.94, 119.44, 118.14, 114.41, 110.99, 110.51, 109.51, 102.29, 79.31, 17.78; HRMS (ESI) calcd for C30H19BrClN9O5 [M+H]+ 702.0461, found 702.0449.

3.4. Insecticidal Activity

The diamondback moth, (Plutella xylostella) was reared in the laboratory for over ten years using vermiculite cultured radish (Raphanus sativusL. var. cuiqing) seedlings. Bioassays were conducted using a leaf-dip method slightly adapted from the methods of Liang et al. [20] and He et al. [21]. Cabbage (Brassica oleracea variant L.) leaves measuring 6 × 6 cm were immersed for 10 s in various concentrations of test compound prepared with distilled water containing 1 g L-1 Triton X-100. Each leaf was left to air dry for 1.5 h and then placed into a Petri dish lined with filter paper. A total of 15 first day fourth instar larvae was introduced into each dish, and three replicates were prepared. Five to seven concentrations of test compound and one control (distilled water with 1 g/L Triton X-100) were examined in each bioassay. Mortality was assessed after 96 h of exposure as individuals that did not move when pushed gently with a brush. The LC50 value was calculated using PoLoPlus 2.0 software (LeOra Software, Petaluma, CA, USA) with data corrected for control mortality by Abbott’s formula [22].

3.5. RyR Radioligand Assay

[3H]Ry (95 Ci/mmol, Perkin-Elmer Life Sciences, Boston, MA, USA) and [3H]Chlo (78 Ci/mmol) were used in RyR binding assays with house fly thorax muscle membranes as reported earlier [15]. Adults emerging from pupae obtained from Benzon Research (Carlisle, PA, USA) were used to collect the thoraces for membrane preparation. The following two buffers were used: (A) 10 μM phenylmethanesulfonyl fluoride, 0.8% bovine serum albumin, and 303 mM sucrose in 20 mM Tris-maleate, pH 7.0; and (B) 0.8 mM CaCl2, 2 mM ATP·Mg2+ salt, and 1.5 M KCl in 10 mM Hepes, pH 7.4. Incubation mixtures were prepared by sequential addition to culture tubes of 200 μL of buffer B, 200 μL of buffer A containing thorax muscle membranes (200 μg protein) and then ethanol (5 μL) containing [3H]Ry or [3H]Chlo to give a final radioligand concentration of 1 nM. Finally, test compounds were added in 5 μL ethanol to give the specified concentrations of Fluor-Cl and Fluor-CN. Following incubation for 2 h at 37 °C, the mixtures were filtered through GF/B filters (Whatman, presoaked in ice-cold washing buffer) immediately after dilution with 5 mL of ice-cold washing buffer (150 mM KCl, 10 mM Hepes, pH 7.4). Each assay tube was further rinsed twice with 5 mL of ice-cold washing buffer, and the rinses were passed through the same filter. The filters were then transferred to scintillation vials containing 10 mL of Safety-Solve (Research Products International Corporation, Mount Prospect, IL, USA) and held overnight in the dark before scintillation counting. Binding data were analyzed and plotted by GraphPad Prism 5.0. All data reported are mean ± standard error for two independent experiments with triplicate samples.

4. Conclusions

The goal to prepare a highly potent and specific fluorescent probe for the insect RyR has been achieved with an anthranilic diamide analog of Chlo containing a hydroxycoumarin substituent. In a validated Musca RyR assay Fluor-Cl is similar to Chlo in potency for stimulating [3H]Ry binding and inhibiting [3H]Chlo binding all in the range of 5–39 nM. The fluorescent probes are less active than Chlo and Cyan as insecticides as expected for compounds with large polar substituents affecting transport and stability. Large species differences are evident in RyR ligand binding with house fly and honeybee quite sensitive [15,16] and therefore possibly preferred RyR sources. Replacing radioligands with fluorescent probes will allow broader use of binding assays without restrictions associated with radioactive materials.

Supplementary Materials

Supplementary materials can be accessed at: https://www.mdpi.com/1420-3049/19/4/4105/s1.

Acknowledgments

This work was supported by the National Natural Science Foundation of China (NNSFC) (No. 20972186) and the National Basic Research Program of China (No. 2010CB126104).

Author Contributions

This research was carried on by all the authors. Yi Wang and Shangzhong Liu designed the theme of the study and carried out the most of compounds synthesis as well as manuscript preparation. Hao Zhang, Kechang Liu, Ruiquan Liu participated in compounds synthesis. Lei Guo and Pei Liang assisted in insecticidal activity tests. Suzhen Qi and John E. Casida were responsible for validated Musca RyR assay, and John E. Casida contributed a lot to manuscript revision.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Rogers, E.F.; Koniuszy, F.R.; Shavel, J., Jr.; Folkers, K. Plant insecticides. I. Ryanodine, a new alkaloid from Ryania speciosa. J. Am. Chem. Soc. 1948, 70, 3086–3088. [Google Scholar] [CrossRef]
  2. Ebbinghaus-Kintscher, U.; Luemmen, P.; Lobitz, N.; Schulte, T.; Funke, C.; Fischer, R.; Masaki, T.; Yasokawa, N.; Tohnishi, M. Phthalic acid diamides activate ryanodine-sensitive Ca2+ release channels in insects. Cell Calcium 2006, 39, 21–33. [Google Scholar] [CrossRef]
  3. Schulte, T.; Ebbinghaus-Kintscher, U.; Luemmen, P.; Fischer, R.; Funke, C. Method for the identification of pesticides. EP 1522855, 13 April 2005. [Google Scholar]
  4. Masaki, T. New insecticide affecting ryanodine receptor, flubendiamide: Biochemical aspects of its action. J. Pestic. Sci. 2006, 31, 484–488. [Google Scholar] [CrossRef]
  5. Masaki, T.; Yasokawa, N.; Tohnishi, M.; Nishimatsu, T.; Tsubata, K.; Inoue, K.; Motoba, K.; Hirooka, T. Flubendiamide, a novel Ca2+ channel modulator, reveals evidence for functional cooperation between Ca2+ pumps and Ca2+ release. Mol. Pharmacol. 2006, 69, 1733–1739. [Google Scholar] [CrossRef]
  6. Luemmen, P.; Ebbinghaus-Kintscher, U.; Funke, C.; Fischer, R.; Masaki, T.; Yasokawa, N.; Tohnishi, M. Phthalic acid diamides activate insect ryanodine receptors. ACS Symp. Ser. 2007, 948, 235–248. [Google Scholar] [CrossRef]
  7. Masaki, T.; Yasokawa, N.; Ebbinghaus-Kintscher, U.; Luemmen, P. Flubendiamide stimulates Ca2+ pump activity coupled to RyR-mediated calcium release in lepidopterous insects. Pestic. Chem. 2007, 137–140. [Google Scholar]
  8. Kato, K.; Kiyonaka, S.; Sawaguchi, Y.; Tohnishi, M.; Masaki, T.; Yasokawa, N.; Mizuno, Y.; Mori, E.; Inoue, K.; Hamachi, I.; et al. Molecular characterization of flubendiamide sensitivity in the lepidopterous ryanodine receptor Ca2+ release channel. Biochemistry 2009, 48, 10342–10352. [Google Scholar] [CrossRef]
  9. Masaki, T.; Yasokawa, N.; Fujioka, S.; Motoba, K.; Tohnishi, M.; Hirooka, T. Quantitative relationship between insecticidal activity and Ca2+ pump stimulation by flubendiamide and its related compounds. J. Pestic. Sci. 2009, 34, 37–42. [Google Scholar] [CrossRef]
  10. Caspar, T.; Cordova, D.; Gutteridge, S.; Rauh, J.J.; Smith, R.M.; Wu, L.; Tao, Y. Cloning and characterization of insect ryanodine receptors and their use for screening for insecticidal compounds. WO 2004027042, 23 September 2003. [Google Scholar]
  11. Cordova, D.; Benner, E.A.; Sacher, M.D.; Rauh, J.J.; Sopa, J.S.; Lahm, G.P.; Selby, T.P.; Stevenson, T.M.; Flexner, L.; Gutteridge, S.; et al. Anthranilic diamides: A new class of insecticides with a novel mode of action, ryanodine receptor activation. Pestic. Biochem. Physiol. 2006, 84, 196–214. [Google Scholar] [CrossRef]
  12. Cordova, D.; Benner, E.A.; Sacher, M.D.; Rauh, J.J.; Sopa, J.S.; Lahm, G.P.; Selby, T.P.; Stevenson, T.M.; Flexner, L.; Caspar, T.; et al. Elucidation of the mode of action of Rynaxypyr, a selective ryanodine receptor activator. In Pesticide Chemistry: Crop Protection, Public Health, Environmental Safety; Ohkawa, H., Miyagawa, H., Lee, P.W., Eds.; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2007; pp. 121–126. [Google Scholar]
  13. Cordova, D.; Benner, E.A.; Sacher, M.D.; Rauh, J.J.; Sopa, J.S.; Lahm, G.P.; Selby, T.P.; Stevenson, T.M.; Flexner, L.; Gutteridge, S.; et al. The novel mode of action of anthranilic diamide insecticides: Ryanodine receptor activation. ACS Symp. Ser. 2007, 948, 223–234. [Google Scholar] [CrossRef]
  14. Lahm, G.P.; Cordova, D.; Barry, J.D. New and selective ryanodine receptor activators for insect control. Bioorg. Med. Chem. 2009, 17, 4127–4133. [Google Scholar] [CrossRef]
  15. Isaacs, A.K.; Qi, S.; Sarpong, R.; Casida, J.E. Insect ryanodine receptor: Distinct but coupled insecticide binding sites for [N-C3H3]chlorantraniliprole, flubendiamide, and [3H]ryanodine. Chem. Res. Toxicol. 2012, 25, 1571–1573. [Google Scholar] [CrossRef]
  16. Qi, S.; Casida, J.E. Species differences in chlorantraniliprole and flubendiamide insecticide binding sites in the ryanodine receptor. Pestic. Biochem. Physiol. 2013, 107, 321–326. [Google Scholar] [CrossRef]
  17. Wang, X.; Wu, S.; Yang, Y.; Wu, Y. Molecular cloning, characterization and mRNA expression of a ryanodine receptor gene from diamondback moth, Plutella xylostella. Pestic. Biochem. Physiol. 2012, 102, 204–212. [Google Scholar] [CrossRef]
  18. Chai, B.; Peng, Y.; Li, H.; Zhang, H.; Liu, C. Synthesis of chlorantraniliprole and its insecticidal activity. Agrochemical 2009, 48, 13–16. [Google Scholar]
  19. Sivakumar, K.; Xie, F.; Cash, B.M.; Long, S.; Barnhill, H.N.; Wang, Q. A fluorogenic 1,3-dipolar cycloaddition reaction of 3-azidocoumarins and acetylenes. Org. Lett. 2004, 6, 4603–4606. [Google Scholar] [CrossRef]
  20. Liang, P.; Gao, X.W.; Zheng, B.Z. Genetic basis of resistance and studies on cross-resistance in a population of diamondback moth, Plutella xylostella (Lepidoptera: Plutellidae). Pest Manag. Sci. 2003, 59, 1232–1236. [Google Scholar] [CrossRef]
  21. He, Y.; Zhao, J.; Zheng, Y.; Desneux, N.; Wu, K. Lethal effect of imidacloprid on the coccinellid predator Serangium japonicum and sublethal effects on predator voracity and on functional response to the whitefly Bemisia tabaci. Ecotoxicology 2012, 21, 1291–1300. [Google Scholar] [CrossRef]
  22. Abbott, W.S. A method of computing the effectiveness of an insecticide. J. Econ. Entomol. 1925, 18, 265–267. [Google Scholar]
  • Sample Availability: Samples of the compounds are available from the authors.

Share and Cite

MDPI and ACS Style

Wang, Y.; Guo, L.; Qi, S.; Zhang, H.; Liu, K.; Liu, R.; Liang, P.; Casida, J.E.; Liu, S. Fluorescent Probes for Insect Ryanodine Receptors: Candidate Anthranilic Diamides. Molecules 2014, 19, 4105-4114. https://doi.org/10.3390/molecules19044105

AMA Style

Wang Y, Guo L, Qi S, Zhang H, Liu K, Liu R, Liang P, Casida JE, Liu S. Fluorescent Probes for Insect Ryanodine Receptors: Candidate Anthranilic Diamides. Molecules. 2014; 19(4):4105-4114. https://doi.org/10.3390/molecules19044105

Chicago/Turabian Style

Wang, Yi, Lei Guo, Suzhen Qi, Hao Zhang, Kechang Liu, Ruiquan Liu, Pei Liang, John E. Casida, and Shangzhong Liu. 2014. "Fluorescent Probes for Insect Ryanodine Receptors: Candidate Anthranilic Diamides" Molecules 19, no. 4: 4105-4114. https://doi.org/10.3390/molecules19044105

Article Metrics

Back to TopTop