Next Article in Journal
Upregulated Palmitoleate and Oleate Production in Escherichia coli Promotes Gentamicin Resistance
Previous Article in Journal
Synthesis of Boronate Affinity-Based Oriented Dummy Template-Imprinted Magnetic Nanomaterials for Rapid and Efficient Solid-Phase Extraction of Ellagic Acid from Food
Previous Article in Special Issue
Synthesis and Anticancer Evaluation of 4-Chloro-2-((5-aryl-1,3,4-oxadiazol-2-yl)amino)phenol Analogues: An Insight into Experimental and Theoretical Studies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Design, Synthesis, Antifungal Activity, and 3D-QSAR Study of Novel Quinoxaline-2-Oxyacetate Hydrazide

Jiangsu Key Laboratory of Pesticide Science, College of Sciences, Nanjing Agricultural University, Nanjing 210095, China
*
Author to whom correspondence should be addressed.
Molecules 2024, 29(11), 2501; https://doi.org/10.3390/molecules29112501
Submission received: 28 April 2024 / Revised: 15 May 2024 / Accepted: 22 May 2024 / Published: 25 May 2024
(This article belongs to the Special Issue Research and Development of Antibacterial and Antitumor Drugs)

Abstract

:
Plant pathogenic fungi pose a major threat to global food security, ecosystem services, and human livelihoods. Effective and broad-spectrum fungicides are needed to combat these pathogens. In this study, a novel antifungal 2-oxyacetate hydrazide quinoxaline scaffold as a simple analogue was designed and synthesized. Their antifungal activities were evaluated against Botrytis cinerea (B. cinerea), Altemaria solani (A. solani), Gibberella zeae (G. zeae), Rhizoctonia solani (R. solani), Colletotrichum orbiculare (C. orbiculare), and Alternaria alternata (A. alternata). These results demonstrated that most compounds exhibited remarkable inhibitory activities and possessed better efficacy than ridylbacterin, such as compound 15 (EC50 = 0.87 μg/mL against G. zeae, EC50 = 1.01 μg/mL against C. orbiculare) and compound 1 (EC50 = 1.54 μg/mL against A. alternata, EC50 = 0.20 μg/mL against R. solani). The 3D-QSAR analysis of quinoxaline-2-oxyacetate hydrazide derivatives has provided new insights into the design and optimization of novel antifungal drug molecules based on quinoxaline.

Graphical Abstract

1. Introduction

Crop pathogens pose a significant threat to food production, with fungicides serving as vital tools for disease control [1,2]. However, prolonged fungicide usage inevitably fosters increased pesticide resistance [3,4]. Therefore, the ongoing development of novel fungicides with distinctive pharmacodynamic frameworks and mechanisms of action is crucial to ensure sustainable agricultural development.
Natural products exhibit diverse structures, encompassing varied biological activities and unique mechanisms of action [5,6]. With excellent environmental compatibility, they serve as crucial sources of lead molecules for drug discovery [7,8]. Quinoxaline, a significant nitrogen-containing natural product, and its derivatives showcase a broad spectrum of biological activities [9,10,11], including antibacterial [12,13], antituberculosis [14], antimalarial [15], antiviral [16], anti-HIV [17], antifungal [18,19], herbicidal [20,21], and bacteriostatic [22] properties. Due to their extensive biological effects, quinoxaline derivatives hold pivotal positions in drug development research [23]. Notable examples include Glecaprevir, utilized in the treatment of chronic hepatitis C, and quinacillin, an antibiotic effective against severe bacterial infections (Figure 1).
In recent years, there has been a growing interest in the synthesis of hydrazides and their agricultural biological activities [24,25,26,27,28,29,30]. Compounds with a hydrazyl formate structure have demonstrated notable biological efficacy. For instance, the broad-spectrum fungicide Miexiuyihao (Figure 1) is a prominent example. Modifying the position of the oxygen element in the ester group has proven to be an effective strategy for enhancing biological activity. In 2021 [31], Chen et al. designed and synthesized quinoline 4-oxyacetate hydrazide derivatives, among which Ac12 exhibited remarkable activity with EC50 values of 0.52 and 0.50 μg/mL against S. sclerotiorum and B. cinerea, surpassing the potency of both commercial fungicides azoxystrobin (both > 30.00 μg/mL) and 8-hydroxyquinoline (2.12 and 5.28 μg/mL). In this study, we focused on the design and synthesis of a novel series of quinoxolinylhydrazide derivatives. These compounds were systematically evaluated for their efficacy against six prominent crop pathogenic fungi through in vitro assays. Subsequently, the most promising compounds were chosen for a further investigation of their in vivo activity. Additionally, we sought to elucidate the primary mechanism of action using scanning electron microscopy (SEM). Furthermore, to gain deeper insights into the structure–activity relationship, a three-dimensional quantitative structure–activity relationship (3D-QSAR) model was established.
Figure 1. Design strategy of title compounds [31].
Figure 1. Design strategy of title compounds [31].
Molecules 29 02501 g001

2. Results and Discussion

Chemistry. The synthetic route of the target compounds is shown in Figure 2. The structures of all compounds were identified by HRMS 1H NMR and 13C NMR spectroscopy (Supplementary Materials).
In Vitro Antifungal Activity Screening. The preliminary results, presented in Table 1, illustrated the significant inhibitory activity of the target compounds against six pathogenic fungi (B. cinerea, A. solani, G. zeae, R. solani, C. orbiculare, and A. alternata) at a concentration of 20.00 μg/mL. Notably, the inhibitory activity against B. cinerea stood out, with compounds 6, 20, and 24 achieving impressive inhibitory rates exceeding 90.0%, at 97.0%, 93.8%, and 93.7%, respectively, surpassing the control drug pyrimethanil at 75.1%. While the inhibitory effect against A. solani was slightly lower, compound 20 still exhibited notable inhibitory activity of 94.9%. Additionally, the inhibitory effect against G. zeae was remarkable, with compounds 1, 2, 7, 15, 16, 18, 26, and 36 displaying inhibitory rates above 90.0%, at 95.8%, 93.1%, 99.7%, 92.2%, 93.6%, 91.4%, 98.9%, and 91.2%, respectively, surpassing the control drug pyrimethanil. The most effective inhibition was observed against R. solani, with 24 compounds exhibiting inhibitory rates above 90.0%, including 8 compounds that completely halted its growth. Furthermore, the inhibitory activity against C. orbiculare was noteworthy, with 6 compounds capable of completely suppressing mycelial growth. For A. alternata, the inhibitory activity was outstanding, with compounds 2 and 7 exhibiting perfect inhibitory rates of 100%, and compounds 1, 3, 6, 16, 26, and 36 demonstrating inhibitory rates above 90.0%. These findings clearly surpass those of the control drug pyrimethanil. The preliminary analysis of structure–activity relationships indicated that compounds 11 through 14 exhibited a relatively low effectiveness, with notably weaker inhibitory effects against R. solani compared to other substituted compounds. While some compounds retained their inhibitory capacity against C. orbiculare, the overall effectiveness was reduced. The activity levels were generally sustained when the quinoxaline benzene ring featured 6,7-dimethyl substitutions. In contrast, chlorine substitution on the quinoxaline benzene ring led to a decrease in overall activity, yet these compounds were more effective against R. solani. Notably, the activity of hydrazide compounds was significantly enhanced with halogen substitutions, but not when substituted with trifluoromethyl. This variation in activity may be attributable not only to the electronic properties of the substituents but also to their size.
The EC50 values of the compounds with excellent antifungal activity were determined (Table 2). Compound 6 showed good activity against B. cinerea, with an EC50 value of 3.31 μg/mL, which was better than the control drug (3.39 μg/mL). Compound 20 exhibited better inhibitory activity against A. solani with an EC50 value of 4.42 μg/mL, superior to carbendazim (5.46 μg/mL). For G. zeae, 8 compounds (1, 2, 7, 15, 16, 18, 26, and 36) demonstrated outstanding inhibitory activity; the EC50 values were all less than 2.00 μg/mL, significantly better than that of the control drug, ridylbacterin (2.20 μg/mL). Notably, compound 15 exhibited the highest activity, achieving an EC50 value of 0.87 μg/mL. For C. orbiculare, compounds 1, 2, 15, 16, 23, 26, 27, and 36 showed the best inhibitory activity with EC50 values of 1.84, 1.32, 1.01, 1.35, 1.36, 1.61, 1.03, and 2.23 μg/mL, respectively, all better than carbendazim (2.32 μg/mL). Compounds 1 (EC50 = 1.54 μg/mL) and 7 (EC50 = 1.99 μg/mL) showed better inhibitory activity against A. alternata, better than the control drug ridylbacterin (EC50 = 2.07 μg/mL). For R. solani, 29 compounds showed excellent antifungal activity, with the EC50 value of target compounds being less than 1 μg/mL. Particularly, the EC50 value of compound 28 was 0.15 μg/mL, lower than the control drug pyrimethanil (EC50 = 0.21 μg/mL).
In Vivo Antifungal Activity. The inhibitory effect of compound 2 on R. solani in rice leaves was determined at a mass concentration of 200.00 μg/mL. As shown in Figure 3, the inhibition rate of compound 2 was 66.1% at this concentration. Compound 6, which exhibited good activity, was selected for a tomato protection experiment, and its activity against B. cinerea was tested on tomato fruits at concentrations of 100.00 and 200.00 μg/mL. As depicted in Figure 3, compound 6 demonstrated significant inhibitory activity at both concentrations, with inhibition reaching 73.3% at 200.00 μg/mL. In vivo experiments demonstrated that compounds 2 and 6 retained some antifungal activity. They hold potential for further research and development.
Scanning Electron Microscopy Observations. Compound 2 was selected for scanning electron microscopy (SEM) to observe changes in the mycelium after treatment. DMSO served as the blank control. As depicted in Figure 4A–C, mycelia treated with DMSO exhibited a smooth, healthy state with a relatively uniform distribution. No folding, atrophy, or breakage of the mycelia were observed. In contrast, the mycelia of R. solani treated with compound 2 showed noticeable pathological changes. Figure 4D,G revealed that the mycelium of R. solani was disorderly distributed, intertwined, and overlapping. Some mycelia exhibited shrinkage, while Figure 4E,F,H,I showed evident pits and depressions, with some mycelia displaying distinctive protrusions that could be irregular stacks of metabolites. Overall, mycelia treated with compound 2 exhibited obvious lesions compared to those treated with DMSO, potentially affecting the direction and state of the mycelia growth by influencing the cell membrane structure of R. solani.
Study of 3D-QSAR Models Against R. solani. To investigate the correlation between the molecular structure of quinoxolinylhydrazine derivatives and their inhibitory effectiveness against R. solani, effective 3D-QSAR models were developed. Then, 20 compounds (24, 6, 12, 13, 1519, 22, 24, 26, 27, 2932, and 34) were randomly selected to be placed in the “training set” and 10 compounds (1, 5, 9, 14, 20, 23, 25, 28, 33, and 25) were selected as the “test set”. These include both the comparative molecular field analysis (CoMFA) model and the comparative molecular similarity indices analysis (CoMSIA) model. Using compound 28 as the template, all compounds were superimposed. Figure 5A shows the energy-minimized structure of compound 28, and Figure 5B displays the compound superposition diagram. Figure 6 illustrates the error and linear relationship between the pEC50 values predicted by the CoMFA and CoMSIA models and the experimental values.
A partial least squares (PLS) analysis was conducted to establish a correlation between the chemical structures and bioactivities of the target compounds. As presented in Table 3, both the CoMFA model (with q2 = 0.843 and r2 = 0.997) and the CoMSIA model (with q2 = 0.845 and r2 = 0.985) exhibited strong robustness and internal predictive power, fulfilling the criteria of q2 > 0.5 and r2 > 0.8 for a reliable forecasting ability. The pEC50 values (−log EC50) were chosen as the molecular activity data for each compound. Table 4 displays the activity prediction results of the CoMFA and CoMSIA models for both the “training set” and “test set” compounds, indicating a good alignment between the predicted and experimental values, with residuals falling within an acceptable error range.
The following three conclusions can be drawn from the CoMFA and CoMSIA equipotential graphs shown in Figure 7. First, as depicted in Figure 7A,D, large green areas around compound 28 suggest that introducing large groups into phenylhy-drazine enhances anti-R. solani activity—for example, EC50 values 1, 2, 4 (R1 = H, R2 = H, R3 = 4-Cl, Br, CF3, EC50 = 0.20, 0.19, 0.17 μg/mL) > 7 (R1 = H, R2 = H, R3 = 4-F, EC50 = 0.65 μg/mL), 15 (R1 = H, R2 = Me, R3 = 4-Cl, EC50 = 0.16 μg/mL) > 16 (R1 = H, R2 = Me, R3 = 4-F, EC50 = 0.36 μg/mL), 27, 28 (R1 = 6-Cl, R2 = H, R3 = 4-Cl, Br, EC50 = 0.18, 0.15 μg/mL) > 26 (R1 = 6-Cl, R2 = H, R3 = 4-F, EC50 = 0.55 μg/mL). Second, as illustrated in Figure 7B,E, a red region at the benzene ring site of compound 28 hydrazide indicates that electron-withdrawing groups at this site improve its antifungal efficacy—for example, EC50 values 1, 2, 7 (R1 = H, R2 = H, R3 = 4-Cl, Br, F, EC50 = 0.20, 0.19, 0.65 μg/mL) > 10 (R1 = H, R2 = H, R3 = 4-OMe, EC50 = 2.22 μg/mL), 15, 16, 17 (R1 = H, R2 = Me, R3 = 4-Cl, F, Br, EC50 = 0.16, 0.36, 0.32 μg/mL) >18, 20 (R1 = H, R2 = Me, R3 = 4-Me, OMe, EC50 = 0.52, 0.54 μg/mL); according to the in vitro activity data, the inhibition rate of 8 on R. solani was only 39.5%, which was also consistent with the rule. Third, also shown in Figure 7B,E, red color blocks at the quinoxaline benzene ring position suggest that electron-absorbing groups here increase the antifungal activity against R. solani—for example, EC50 values 26, 27, 28 (R1 = 6-Cl, R2 = H, R3 = 4-F, Cl, Br, EC50 = 0.55, 0.18, 0.15 μg/mL) > 1, 2, 7 (R1 = H, R2 = H, R3 = 4-Cl, Br, F, EC50 = 0.20, 0.19, 0.65 μg/mL), 28 (R1 = 6-Cl, R2 = H, R3 = 4-Br, EC50 = 0.15 μg/mL) > 22 (R1 = 6,7-Me, R2 = H, R3 = 4-Br, EC50 = 0.47 μg/mL). It can be seen from the preliminary screening data of in vitro activity that 25 and 35 could completely inhibit the growth of R. solani, and the inhibition rate of 24 was 92.7%. Figure 7 shows that yellow, green, red, and blue color blocks are concentrated on the benzene ring, indicating that various substituents on the benzene ring significantly influence anti-R. solani activity. This observation aligns with the results of activity tests. Generally, the addition of electron-withdrawing groups on the benzene ring enhances anti-R. solani activity, particularly when F, Cl, and Br are introduced at the fourth position, significantly boosting the activity. Chlorine substitution on the quinoxaline ring also enhanced the inhibitory activity against R. solani, although the improvement was modest. This corresponds with the small red color block on quinoxaline in Figure 7B,E, supporting the findings from the 3D-QSAR analysis.

3. Materials and Methods

Instruments and Chemical Reagents. The reagents and solvents utilized are commercially available and do not necessitate additional purification. 1H NMR and 13C NMR were measured by a BRUKER-400 NMR instrument (Bruker Corporation, Rheinstetten, Germany) with CDCl3, or DMSO-d6 as the solvent and tetramethylsilane (TMS) as the internal standard. High-resolution mass spectra (HRMS) data were collected with a Triple TOF 5600 plus LCMS spectrometer (AB Sciex, Framingham, MA, USA). The melting points (mp) were determined using the Buchi M-560 Melting Point Apparatus (BUCHI, Flawil, Switzerland).
General Method of Synthesis. The synthetic route of the compounds is shown in the Figure 2. The synthesis methods of intermediates b and c refer to the reported methods [18,32].
General Synthesis method for the Intermediate d. Intermediate c (5.00 mmol) and Caesium Carbonate (20.00 mmol) were added to 20.00 mL DMSO. We added ethyl glycolate (7.50 mmol) slowly. The reaction was stirred at 80 °C. The reaction was monitored using thin-layer chromatography (TLC) until the reaction was complete. We poured the mixture into cold water and extracted it by ethyl acetate. The organic phase was dried over Na2SO4. Subsequently, the solvent was evaporated to yield the crude product, which underwent purification via column chromatography.
General Synthesis method for the Intermediate e. Intermediate d (5.00 mmol) was added to 20.00 mL 1.00 M NaOH. The reaction was stirred at room temperature and monitored using TLC until the reaction was complete. Then, we adjusted the mixture to acid with 10% HCl, and white solid precipitates formed. The intermediate e was obtained by filtration.
General Synthesis method for Target Compounds 136. Intermediate e (1.00 mmol) and 2-(1H-Benzotriazole-1-yl)-1,1,3,3-tetramethyluronium tetrafluoroborate (TBTU, 1.20 mmol) were added to 4.00 mL CH3CN. Et3N (0.50 mL) was added. The reaction was stirred at room temperature and monitored using TLC until the reaction was complete. The solvent was evaporated to yield the crude product, which underwent purification via column chromatography.
In Vitro Antifungal Activity Screening. The antifungal activity of the target compounds against six pathogenic fungi was tested by mycelium growth method [33] (more details of the target compounds for the antifungal activities test procedure are shown in the Supplementary Materials), including B. Cinerea, A. Solani, G. zeae, R. solani, C. Orbiculare, and A. Alternata. The compounds dissolved in DMSO were subsequently introduced into a quantitative PDA (90.00 g glucose, 1000.00 g potato, 100.00 g agar, and 5.00 L water) medium to create a medicated plate, achieving a final concentration of 20.00 μg/mL. A plate was prepared with equal-volume DMSO as a blank control. All samples were incubated at 25 °C for 2–5 days, and each plate was assessed using the cross-streak method. Mean values and standard deviations (unit: cm) were computed, along with the inhibition rate following drug treatment. The calculation method is referred to as the reported method [34,35].
In Vivo Antifungal Activity. We prepared 500.00 mL of 1% Tsum 80 solution, weighed 10.00 mg of compound 2, and dissolved it in 250.00 μL of DMSO. We used 1% Tsum 80 solution to bring the volume up to 50.00 mL, stirred well, and set it aside. In this case, the liquid concentration was 200.00 μg/mL. The blank control was prepared using 250.00 μL of DMSO and 1% Tsum 80 solution. We used the same method to prepare other required medications. In vivo antifungal activity was tested using reported methods [36,37,38].
Choose compounds exhibiting outstanding antifungal inhibition rates for in vitro EC50 determination. Five concentration gradients of 10.00, 5.00, 2.50, 1.25, and 0.625 μg/mL were established. We assess the inhibition rates of each concentration using the aforementioned method. We utilize the DPS data processing system to calculate the corresponding EC50 for each compound.
Scanning Electron Microscopy Observations. According to the above method, 2 plate containing drug concentration of 10.00 μg/mL was prepared, and blank control was set. R. solani was inoculated on the medicated plate and cultured for 36 h. The cake at the edge of the hypha was taken with a hole punch and fixed with phosphate buffer solution of glutaraldehyde. The mycelium was observed by scanning electron microscopy (Hitachi SU8010, Hitachi Co., Tokyo, Japan) with resolutions of 50.00 µm, 20.00 µm, and 10.00 µm.
Research of 3D-QSAR. Sybyl X-2.0 (Tripos, El Cerrito, CA, USA) was used to build 3D-QSAR models. The minimization of energy for the target molecular conformations was carried out utilizing the Tripos force field, along with the Gasteiger–Hückel method (with 10,000 iterations and 0.005 kcal mol-1 Å-1 convergence gradient), ensuring precision and thoroughness in the energy minimization. The characteristics of the CoMFA and CoMSIA models’ fields were determined through a thorough analysis employing partial least squares (PLS) regression.

4. Conclusions

In summary, a series of quinoxolinylhydrazide derivatives were designed and synthesized. Most of the compounds exhibited notable antifungal activity, with compound 1 showing superior inhibitory effects against six pathogenic fungi. The activity against R. solani was particularly impressive, with 29 compounds exhibiting EC50 values below 1.00 μg/mL, especially compound 28, which had an EC50 of 0.15 μg/mL. In vivo antifungal activity experiments confirmed that compounds 2 and 6 maintained effectiveness within living organisms. Scanning electron microscopy results suggested that compound 2 was a potential inhibitor of R. solani. Reliable 3D-QSAR models targeting R. solani were successfully developed. The results of these experiments will help identify new potential fungicides.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules29112501/s1, Table S1. Antifungal EC50 values of target compounds against phytopathogenic fungia; Scheme S1. Keto-enol tautomerization of compound 3.

Author Contributions

Conceptualization, P.T. and W.Z.; experiment execution and data analysis, P.T., Y.L., R.F., Y.Z. and P.D.; writing—original draft preparation, P.T.; writing—review and editing, P.T. and P.D. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Key Research and Development Program of China (2021YFD1700102).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data are contained within the article and Supplementary Materials.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Li, L.; Zhu, X.-M.; Zhang, Y.-R.; Cai, Y.-Y.; Wang, J.-Y.; Liu, M.-Y.; Wang, J.-Y.; Bao, J.-D.; Lin, F.-C. Research on the Molecular Interaction Mechanism between Plants and Pathogenic Fungi. Int. J. Mol. Sci. 2022, 23, 4658. [Google Scholar] [CrossRef] [PubMed]
  2. Wu, Y.-L.; Han, L.-J.; Wu, X.-M.; Jiang, W.; Liao, H.; Xu, Z.; Pan, C.-P. Trends and perspectives on general Pesticide analytical chemistry. Adv. Agrochem. 2022, 1, 113–124. [Google Scholar] [CrossRef]
  3. Strange, R.N.; Scott, P.R. Plant Disease: A Threat to Global Food Security. Annu. Rev. Phytopathol. 2005, 43, 83–116. [Google Scholar] [CrossRef] [PubMed]
  4. Wang, X.; Fu, X.; Chen, M.; Wang, A.; Yan, J.; Mei, Y.; Wang, M.; Yang, C. Novel 1,3,5-thiadiazine-2-thione derivatives containing a hydrazide moiety: Design, synthesis and bioactive evaluation against phytopathogenic fungi in vitro and in vivo. Chin. Chem. Lett. 2019, 30, 1419–1422. [Google Scholar] [CrossRef]
  5. Swain, T. Secondary Compounds as Protective Agents. Annu. Rev. Plant Biol. 1977, 28, 479–501. [Google Scholar] [CrossRef]
  6. Zhang, P.; Duan, C.-B.; Jin, B.; Ali, A.S.; Han, X.; Zhang, H.; Zhang, M.-Z.; Zhang, W.-H.; Gu, Y.-C. Recent advances in the natural products-based lead discovery for new agrochemicals. Adv. Agrochem. 2023, 2, 324–339. [Google Scholar] [CrossRef]
  7. Arif, T.; Bhosale, J.D.; Kumar, N.; Mandal, T.K.; Bendre, R.S.; Lavekar, G.S.; Dabur, R. Natural products—Antifungal agents derived from plants. J. Asian Nat. Prod. Res. 2009, 11, 621–638. [Google Scholar] [CrossRef] [PubMed]
  8. Roberts, D.P.; Mattoo, A.K. Sustainable Agriculture—Enhancing Environmental Benefits, Food Nutritional Quality and Building Crop Resilience to Abiotic and Biotic Stresses. Agriculture 2018, 8, 8. [Google Scholar] [CrossRef]
  9. Dhameliya, T.M.; Chudasma, S.J.; Patel, T.M.; Dave, B.P. A review on synthetic account of 1,2,4-oxadiazoles as anti-infective agents. Mol. Divers. 2022, 26, 2967–2980. [Google Scholar] [CrossRef]
  10. Xu, H.; Fan, L.-L. Synthesis and antifungal activities of novel 5,6-dihydro-indolo[1,2-a]quinoxaline derivatives. Eur. J. Med. Chem. 2011, 46, 1919–1925. [Google Scholar] [CrossRef]
  11. Jiang, Y.; Zhu, C.-H.; Xia, Z.-H.; Zhao, H.-Q. Quinoxalinone-1,2,3-triazole derivatives as potential antifungal agents for plant anthrax disease: Design, synthesis, antifungal activity and SAR study. Adv. Agrochem. 2023. [Google Scholar] [CrossRef]
  12. El-Zahabi, H.S.A. Synthesis, Characterization, and Biological Evaluation of Some Novel Quinoxaline Derivatives as Antiviral Agents. Arch. Pharm. 2017, 350, 1700028–1700040. [Google Scholar] [CrossRef] [PubMed]
  13. Zhao, Y.; Cheng, G.; Hao, H.; Pan, Y.; Liu, Z.; Dai, M.; Yuan, Z. In vitro antimicrobial activities of animal-used quinoxaline 1,4-di-N-oxides against mycobacteria, mycoplasma and fungi. BMC Vet. Res. 2016, 12, 186–198. [Google Scholar] [CrossRef]
  14. Moreno, E.; Ancizu, S.; Pérez-Silanes, S.; Torres, E.; Aldana, I.; Monge, A. Synthesis and antimycobacterial activity of new quinoxaline-2-carboxamide 1,4-di-N-oxide derivatives. Eur. J. Med. Chem. 2010, 45, 4418–4426. [Google Scholar] [CrossRef]
  15. Shekhar, A.C.; Rao, P.S.; Narsaiah, B.; Allanki, A.D.; Sijwali, P.S. Emergence of pyrido quinoxalines as new family of antimalarial agents. Eur. J. Med. Chem. 2014, 77, 280–287. [Google Scholar] [CrossRef]
  16. Wilhelmsson, L.M.; Kingi, N.; Bergman, J. Interactions of Antiviral Indolo[2,3-b]quinoxaline Derivatives with DNA. J. Med. Chem. 2008, 51, 7744–7750. [Google Scholar] [CrossRef]
  17. De Clercq, E. Toward Improved Anti-HIV Chemotherapy: Therapeutic Strategies for Intervention with HIV Infections. J. Med. Chem. 1995, 38, 2491–2517. [Google Scholar] [CrossRef] [PubMed]
  18. Borah, B.; Chowhan, L.R. Recent advances in the transition-metal-free synthesis of quinoxalines. RSC Adv. 2021, 11, 37325–37353. [Google Scholar] [CrossRef]
  19. Montana, M.; Mathias, F.; Terme, T.; Vanelle, P. Antitumoral activity of quinoxaline derivatives: A systematic review. Eur. J. Med. Chem. 2019, 163, 136–147. [Google Scholar] [CrossRef]
  20. Pereira, J.A.; Pessoa, A.M.; Cordeiro, M.N.; Fernandes, R.; Prudencio, C.; Noronha, J.P.; Vieira, M. Quinoxaline, its derivatives and applications: A State of the Art review. Eur. J. Med. Chem. 2015, 97, 664–672. [Google Scholar] [CrossRef]
  21. Tariq, S.; Alam, O.; Synthesis, M.A. Anti-inflammatory, p38α MAP kinase inhibitory activities and molecular docking studies of quinoxaline derivatives containing triazole moiety. Bioorg. Chem. 2018, 76, 343–358. [Google Scholar] [CrossRef] [PubMed]
  22. Hu, Y.; Wang, K.; MacMillan, J.B. Hunanamycin A, an Antibiotic from a Marine-Derived Bacillus hunanensis. Org. Lett. 2013, 15, 390–393. [Google Scholar] [CrossRef] [PubMed]
  23. Xia, R.; Guo, T.; He, J.; Chen, M.; Su, S.; Jiang, S.; Tang, X.; Chen, Y.; Xue, W. Antimicrobial evaluation and action mechanism of chalcone derivatives containing quinoxaline moiety. Monatsh. Chem. 2019, 150, 1325–1334. [Google Scholar] [CrossRef]
  24. Chang, J.; Liu, Y.; Zhang, T.; Chen, Z.; Fang, H.; Hua, X. A Comprehensive Investigation of Hydrazide and Its Derived Structures in the Agricultural Fungicidal Field. J. Agric. Food Chem. 2023, 71, 8297–8316. [Google Scholar] [CrossRef] [PubMed]
  25. Niu, J.-B.; Quan, C.-H.; Liu, Y.; Yu, G.-X.; Yang, J.-J.; Li, Y.-R.; Zhang, Y.-B.; Qi, Y.-Q.; Song, J.; Jin, C.-Y.; et al. Discovery of N-aryl sulphonamide-quinazoline derivatives as anti-gastric cancer agents in vitro and in vivo via activating the Hippo signalling pathway. J. Enzym. Inhib. Med. Chem. 2021, 36, 1715–1731. [Google Scholar] [CrossRef]
  26. Sharma, P.C.; Sharma, D.; Sharma, A.; Saini, N.; Goyal, R.; Ola, M.; Chawla, R.; Thakur, V.K. Hydrazone comprising compounds as promising anti-infective agents: Chemistry and structure-property relationship. Mater. Today Chem. 2020, 18, 100349–100369. [Google Scholar] [CrossRef]
  27. Stierli, D.; Eberle, M.; Lamberth, C.; Jacob, O.; Balmer, D.; Gulder, T. Quarternary α-cyanobenzylsulfonamides: A new subclass of CAA fungicides with excellent anti-Oomycetes activity. Bioorg. Med. Chem. 2021, 30, 115965. [Google Scholar] [CrossRef] [PubMed]
  28. Wang, Z.-W.; Zhao, L.-X.; Ma, P.; Ye, T.; Fu, Y.; Ye, F. Fragments recombination, design, synthesis, safener activity and CoMFA model of novel substituted dichloroacetylphenyl sulfonamide derivatives. Pest Manag. Sci. 2021, 77, 1724–1738. [Google Scholar] [CrossRef] [PubMed]
  29. Chen, Z.; Fang, H.; Chang, J.; Zhang, T.; Cui, Y.; Zhang, L.; Sui, J.; Ma, Q.; Su, P.; Wang, J.; et al. Natural Alkaloid Waltherione F-Derived Hydrazide Compounds Evaluated in an Agricultural Fungicidal Field. J. Agric. Food Chem. 2023, 71, 12333–12345. [Google Scholar] [CrossRef]
  30. Yao, X.; Zhang, R.; Lv, B.; Wang, W.-W.; Liu, Z.; Hu, Z.; Design, D.L. Synthesis and biological evaluation of thiazole and imidazo[1,2-a]pyridine derivatives containing a hydrazone substructure as potential agrochemicals. Adv. Agrochem. 2023, 2, 154–162. [Google Scholar] [CrossRef]
  31. Chen, Y.-J.; Ma, K.-Y.; Du, S.-S.; Zhang, Z.-J.; Wu, T.-L.; Sun, Y.; Liu, Y.-Q.; Yin, X.-D.; Zhou, R.; Yan, Y.-F.; et al. Antifungal Exploration of Quinoline Derivatives against Phytopathogenic Fungi Inspired by Quinine Alkaloids. J. Agric. Food Chem. 2021, 69, 12156–12170. [Google Scholar] [CrossRef] [PubMed]
  32. Kumar, A.; Dhameliya, T.M.; Sharma, K.; Patel, K.A.; Hirani, R.V.; Bhatt, A.J. Sustainable approaches towards the synthesis of quinoxalines: An update. J. Mol. Struct. 2022, 1259, 132732–132747. [Google Scholar] [CrossRef]
  33. Mao, X.; Wu, Z.; Bi, C.; Wang, J.; Zhao, F.; Gao, J.; Hou, Y.; Zhou, M. Molecular and Biochemical Characterization of Pydiflumetofen-Resistant Mutants of Didymella bryoniae. J. Agric. Food Chem. 2020, 68, 9120–9130. [Google Scholar] [CrossRef] [PubMed]
  34. Yang, Y.D.; He, Y.H.; Ma, K.Y.; Li, H.; Zhang, Z.J.; Sun, Y.; Wang, Y.L.; Hu, G.F.; Wang, R.X.; Liu, Y.Q. Design and Discovery of Novel Antifungal Quinoline Derivatives with Acylhydrazide as a Promising Pharmacophore. J. Agric. Food Chem. 2021, 69, 8347–8357. [Google Scholar] [CrossRef] [PubMed]
  35. Wang, X.; Dai, Z.C.; Chen, Y.F.; Cao, L.L.; Yan, W.; Li, S.K.; Wang, J.X.; Zhang, Z.G. Synthesis of 1,2,3-triazole hydrazide derivatives exhibiting anti-phytopathogenic activity. Eur. J. Med. Chem. 2017, 126, 171–182. [Google Scholar] [CrossRef] [PubMed]
  36. Wang, X.; Wang, A.; Qiu, L.; Chen, M.; Lu, A.; Li, G.; Yang, C.; Xue, W. Expedient Discovery for Novel Antifungal Leads Targeting Succinate Dehydrogenase: Pyrazole-4-formylhydrazide Derivatives Bearing a Diphenyl Ether Fragment. J. Agric. Food Chem. 2020, 68, 14426–14437. [Google Scholar] [CrossRef]
  37. Yang, C.; Sun, S.; Li, W.; Mao, Y.; Wang, Q.; Duan, Y.; Csuk, R.; Li, S. Bioactivity-Guided Subtraction of MIQOX for Easily Available Isoquinoline Hydrazides as Novel Antifungal Candidates. J. Agric. Food Chem. 2023, 71, 11341–11349. [Google Scholar] [CrossRef]
  38. Zhi, X.Y.; Zhang, Y.; Li, Y.F.; Liu, Y.; Niu, W.P.; Li, Y.; Zhang, C.R.; Cao, H.; Hao, X.J.; Yang, C. Discovery of Natural Sesquiterpene Lactone 1-O-Acetylbritannilactone Analogues Bearing Oxadiazole, Triazole, or Imidazole Scaffolds for the Development of New Fungicidal Candidates. J. Agric. Food Chem. 2023, 71, 11680–11691. [Google Scholar] [CrossRef]
Figure 2. The synthetic route of the target compounds. Reagents and conditions. (i) H2O, SDS; (ii) POCl3, reflux; (iii) CsCO3, DMSO, 80 °C, 5 h; (iv) NaOH, H2O, rt; HCl, H2O, rt; and (v) TBTU, Et3N, CH3CN. (ae) are reaction intermediate.
Figure 2. The synthetic route of the target compounds. Reagents and conditions. (i) H2O, SDS; (ii) POCl3, reflux; (iii) CsCO3, DMSO, 80 °C, 5 h; (iv) NaOH, H2O, rt; HCl, H2O, rt; and (v) TBTU, Et3N, CH3CN. (ae) are reaction intermediate.
Molecules 29 02501 g002
Figure 3. The in vivo preventive effects of compound 2 against R. solani and compound 6 against B. cinerea.
Figure 3. The in vivo preventive effects of compound 2 against R. solani and compound 6 against B. cinerea.
Molecules 29 02501 g003
Figure 4. SEM images of R. solani hyphae treated by DMSO (AC) and 2 (DI).
Figure 4. SEM images of R. solani hyphae treated by DMSO (AC) and 2 (DI).
Molecules 29 02501 g004
Figure 5. Template compound 28 (A) and composite renderings (B).
Figure 5. Template compound 28 (A) and composite renderings (B).
Molecules 29 02501 g005
Figure 6. Experimental and predicted values of CoMFA (A) and CoMSIA (B) models pEC50.
Figure 6. Experimental and predicted values of CoMFA (A) and CoMSIA (B) models pEC50.
Molecules 29 02501 g006
Figure 7. CoMFA (A,B) contour maps and CoMSIA (CF) contour maps with compound 28 shown inside fields. Green and yellow modules indicate regions where steric bulk would increase and reduce anti-R. solani activity; blue and red modules indicate regions where electropositive and electronegative groups would increase anti-R. solani activity; yellow and grey modules indicate regions where hydrophobicity and hydrophilicity would increase anti-R. solani activity; magenta and red modules indicate regions where H-bond acceptor groups would increase and reduce the anti-R. solani activity, meanwhile, cyan and purple modules indicate regions where H-donor groups would increase and reduce anti-R. solani activity.
Figure 7. CoMFA (A,B) contour maps and CoMSIA (CF) contour maps with compound 28 shown inside fields. Green and yellow modules indicate regions where steric bulk would increase and reduce anti-R. solani activity; blue and red modules indicate regions where electropositive and electronegative groups would increase anti-R. solani activity; yellow and grey modules indicate regions where hydrophobicity and hydrophilicity would increase anti-R. solani activity; magenta and red modules indicate regions where H-bond acceptor groups would increase and reduce the anti-R. solani activity, meanwhile, cyan and purple modules indicate regions where H-donor groups would increase and reduce anti-R. solani activity.
Molecules 29 02501 g007
Table 1. Inhibitory activities of target compounds 136 against phytopathogenic fungi a.
Table 1. Inhibitory activities of target compounds 136 against phytopathogenic fungi a.
Compd.B. cinreaA. solaniG. zeaeR. solaniC. orbiculareA. alternata
186.1 ± 1.489.5 ± 1.295.8 ± 0.891.9 ± 0.910096.4 ± 0.5
275.9 ± 0.548.1 ± 1.393.1 ± 0.694.4 ± 0.897.7 ± 0.5100
388.8 ± 1.679.3 ± 0.759.7 ± 2.297.2 ± 0.710091.2 ± 0.7
474.4 ± 1.666.4 ± 2.465.0 ± 1.277.4 ± 1.387.3 ± 1.185.0 ± 1.5
539.6 ± 3.040.7 ± 0.785.3 ± 0.696.4 ± 0.781.9 ± 0.660.9 ± 0.5
697.0 ± 0.680.6 ± 026.3 ± 0.770.2 ± 0.810095.3 ± 2
780.8 ± 1.785.0 ± 0.799.7 ± 0.610051.3 ± 1.0100
810.2 ± 0.528.8 ± 1.0<1039.5 ± 0.8<10<10
985.6 ± 1.473.7 ± 1.229.2 ± 1.397.2 ± 0.756.2 ± 1.146.9 ± 0.5
1021.9 ± 1.147.9 ± 0.731.9 ± 2.081.9 ± 2.150 ± 0.641.1 ± 0.7
1114.7 ± 1.658.6 ± 1.012.8 ± 1.281.0 ± 0.783.5 ± 0.560.4 ± 0.7
1252.2 ± 1.056.5 ± 1.925.6 ± 1.273.9 ± 2.250.1 ± 0.772.0 ± 0.6
1356.7 ± 1.042.1 ± 0.725.8 ± 1.975.0 ± 0.835.6 ± 0.673.6 ± 0.8
1428.0 ± 1.260.7 ± 0.723.9 ± 0.879.6 ± 1.234.4 ± 2.677.1 ± 0.7
1580.6 ± 1.76.4 ± 0.792.2 ± 1.298.0 ± 0.791.7 ± 0.663.2 ± 0.7
1670.6 ± 1.162.1 ± 0.793.6 ± 0.694.4 ± 0.895.6 ± 0.697.9 ± 0.7
1774.6 ± 1.766.4 ± 0.784.7 ± 1.196.4 ± 0.760.1 ± 2.948.1 ± 0.7
1881.6 ± 1.465.1 ± 0.891.4 ± 0.698.0 ± 1.394.8 ± 0.787.8 ± 2.4
1920.4 ± 0.733.6 ± 0.742.2 ± 2.373.4 ± 0.818.8 ± 0.723.1 ± 0.7
2093.8 ± 0.794.9 ± 1.448.1 ± 4.398.4 ± 0.710063.8 ± 3.4
2155.5 ± 2.035.0 ± 0.745.0 ± 1.073.4 ± 0.847.3 ± 0.533.1 ± 0.5
2260.7 ± 2.476.4 ± 0.730.8 ± 0.897.2 ± 0.734.6 ± 0.639.6 ± 0.7
2375.4 ± 0.769.1 ± 0.780.8 ± 1.697.6 ± 0.810032.4 ± 0.7
2493.7 ± 0.684.9 ± 1.225.8 ± 1.392.7 ± 0.767.2 ± 3.474.3 ± 0.6
2541.5 ± 1.119.3 ± 0.715.8 ± 2.510049.4 ± 0.650.7 ± 0.7
2676.6 ± 1.162.1 ± 0.798.9 ± 0.810010091.9 ± 0.5
2769.9 ± 2.646.4 ± 0.789.7 ± 2.295.2 ± 0.990.6 ± 0.656.3 ± 0.7
2844.5 ± 2.039.3 ± 0.742.5 ± 1.678.0 ± 2.279.4 ± 0.645.8 ± 0.8
29<1020.0 ± 1.0<1010014 ± 0.5<10
3061.7 ± 0.720.7 ± 0.733.1 ± 0.610039.4 ± 0.654.9 ± 0.7
31<1023.7 ± 4.427.5 ± 4.496.8 ± 1.118.2 ± 1.122.2 ± 2.3
3246.6 ± 1.154.6 ± 1.178.9 ± 2.683.9 ± 0.970.6 ± 1.460.2 ± 2.6
3355.2 ± 3.253.3 ± 1.468.9 ± 1.992.5 ± 6.370.4 ± 1.348.4 ± 1.8
3450.5 ± 1.570.7 ± 0.685.5 ± 1.110065.9 ± 1.645.1 ± 1.4
3514.1 ± 2.041.1 ± 0.824.8 ± 1.110039.6 ± 1.729.3 ± 1.6
3657.8 ± 3.386.1 ± 1.291.2 ± 0.510010092.0 ± 1.9
pyrimethanil75.1 ± 1.043.1 ± 0.732.6 ± 1.889.4 ± 0.715.2 ± 1.127.2 ± 1.5
a Average of three replicates.
Table 2. Antifungal EC50 values of target compounds against phytopathogenic fungi a.
Table 2. Antifungal EC50 values of target compounds against phytopathogenic fungi a.
PathogenCompd.EC50PathogenCompd.EC50
B. Cinerea63.31 ± 0.18R. solani10.20 ± 0.07
204.36 ± 0.1020.19 ± 0.04
264.90 ± 0.0530.58 ± 0.03
pyrimethanil3.39 ± 0.2240.17 ± 0.11
A. solani204.42 ± 0.0950.26 ± 0.07
carbendazim5.46 ± 0.1460.58 ± 0.16
G. zeae10.94 ± 0.0370.65 ± 0.08
21.22 ± 0.0690.84 ± 0.20
71.21 ± 0.02102.22 ± 0.06
150.87 ± 0.03111.41 ± 0.11
161.17 ± 0.08121.05 ± 0.04
181.77 ± 0.04130.20 ± 0.01
281.27 ± 0.06140.39 ± 0.08
381.54 ± 0.09150.16 ± 0.04
pyrimethanil2.20 ± 0.11160.36 ± 0.12
C. orbiculare11.84 ± 0.07170.32 ± 0.23
21.32 ± 0.06180.52 ± 0.03
33.35 ± 0.22190.33 ± 0.11
68.39 ± 0.15200.54 ± 0.14
151.01 ± 0.11210.15 ± 0.09
161.35 ± 0.07220.47 ± 0.04
183.84 ± 0.10230.54 ± 0.14
203.86 ± 0.13240.66 ± 0.11
231.36 ± 0.14250.68 ± 0.05
261.61 ± 0.12260.55 ± 0.05
271.03 ± 0.11270.18 ± 0.04
362.23 ± 0.08280.15 ± 0.08
carbendazim2.32 ± 0.11291.21 ± 0.11
A. alternata11.54 ± 0.12301.05 ± 0.09
210.75 ± 0.05310.80 ± 0.11
37.35 ± 0.21320.23 ± 0.05
612.09 ± 0.05330.38 ± 0.07
71.99 ± 0.08340.36 ± 0.05
164.82 ± 0.25350.55 ± 0.23
282.85 ± 0.06360.50 ± 0.15
pyrimethanil2.07 ± 0.15pyrimethanil0.21 ± 0.10
a Average of three replicates.
Table 3. Statistical parameters of CoMFA and CoMSIA models.
Table 3. Statistical parameters of CoMFA and CoMSIA models.
Statistical ParameterCoMFACoMSIAValidation Criteria
q2 a0.8430.845>0.5
r2 b0.9970.985>0.8
s c0.0250.038
F d144.141125.981
ONC e147
a Cross-validated correlation. b Non-cross-validated correlation. c Standard error of estimate. d F-test value. e Optimum number of components.
Table 4. Actual and predicted pEC50 values of CoMFA and CoMSIA a.
Table 4. Actual and predicted pEC50 values of CoMFA and CoMSIA a.
Compd.EC50pEC50CoMFACoMSIA
Pred pEC50 bResidualPred pEC50 bResidual
1 * 0.20 ± 0.076.3116.2950.0166.324−0.013
20.19 ± 0.046.3436.382−0.0396.375−0.032
30.58 ± 0.035.7285.7270.0015.759−0.031
40.17 ± 0.116.3226.2710.0516.341−0.019
5 * 0.26 ± 0.076.0716.376−0.3056.29−0.219
60.58 ± 0.165.7296.193−0.4645.6970.032
9 * 2.22 ± 0.065.6456.209−0.5645.60.045
121.05 ± 0.045.6215.933−0.3125.973−0.352
130.20 ± 0.016.0156.022−0.0076.034−0.019
14 * 0.39 ± 0.086.0566.0490.0076.0360.02
150.12 ± 0.046.0985.9770.1215.9670.131
160.36 ± 0.125.9745.9670.0075.978−0.004
170.32 ± 0.236.0046.013−0.0095.9990.005
180.52 ± 0.035.7925.800−0.0085.7870.005
190.33 ± 0.116.0536.0520.0016.061−0.008
20 * 0.54 ± 0.145.785.770.015.7740.006
220.47 ± 0.045.9335.7140.2195.620.313
23 * 0.54 ± 0.145.7985.805−0.0075.819−0.021
240.66 ± 0.115.7145.7060.0085.7070.007
25 * 0.68 ± 0.055.7075.72−0.0135.6730.034
260.55 ± 0.055.7966.184−0.3886.253−0.457
270.18 ± 0.046.3086.3010.0076.322−0.014
28 * 0.15 ± 0.086.4226.3910.0316.3560.066
291.21 ± 0.115.5275.644−0.1175.561−0.034
301.05 ± 0.095.5405.554−0.0145.62−0.08
310.80 ± 0.115.6935.6890.0045.6880.005
320.23 ± 0.056.2426.0480.1945.9920.25
33 * 0.38 ± 0.076.0596.33−0.2716.0450.014
340.36 ± 0.056.0926.33−0.2386.317−0.225
35 * 0.55 ± 0.235.8395.84−0.0015.846−0.007
a Average of three replicates. b pEC50 = −log (EC50); the unit of EC50 value needs to be converted into mol/L. * Sample of the testing set.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Teng, P.; Li, Y.; Fang, R.; Zhu, Y.; Dai, P.; Zhang, W. Design, Synthesis, Antifungal Activity, and 3D-QSAR Study of Novel Quinoxaline-2-Oxyacetate Hydrazide. Molecules 2024, 29, 2501. https://doi.org/10.3390/molecules29112501

AMA Style

Teng P, Li Y, Fang R, Zhu Y, Dai P, Zhang W. Design, Synthesis, Antifungal Activity, and 3D-QSAR Study of Novel Quinoxaline-2-Oxyacetate Hydrazide. Molecules. 2024; 29(11):2501. https://doi.org/10.3390/molecules29112501

Chicago/Turabian Style

Teng, Peng, Yufei Li, Ruoyu Fang, Yuchuan Zhu, Peng Dai, and Weihua Zhang. 2024. "Design, Synthesis, Antifungal Activity, and 3D-QSAR Study of Novel Quinoxaline-2-Oxyacetate Hydrazide" Molecules 29, no. 11: 2501. https://doi.org/10.3390/molecules29112501

Article Metrics

Back to TopTop