Cryo-EM structure of the yeast Saccharomyces cerevisiae SDH provides a template for eco-friendly fungicide discovery
Zhi-Wen Li
Yuan-Hui Huang
Ge Wei
Zong-Wei Lu
Yu-Xia Wang
Guang-Rui Cui
Jun-Ya Wang
Xin-He Yu
Yi-Xuan Fu
Er-Di Fan
Qiong-You Wu
Xiao-Lei Zhu
Ying Ye
Guang-Fu Yang
Corresponding author.
Contributed equally.
Received 2025 Mar 4; Accepted 2025 Sep 4; Collection date 2025.
Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
Abstract
Succinate dehydrogenase (SDH) is a key fungicidal target, but rational inhibitors design has been impeded by the lack of fungal SDH structure. Here, we show the cryo-EM structure of SDH from Saccharomyces cerevisiae (ScSDH) in apo (3.36 Å) and ubiquinone-1-bound (3.25 Å) states, revealing subunits architecture and quinone-binding sites (Qp). ScSDH is classified as a heme-deficient type-D SDH, utilizing conserved redox centers (FAD, [2Fe-2S], [4Fe-4S] and [3Fe-4S] clusters) for electron transfer. A 3.23 Å structure with pydiflumetofen (PYD) identified critical interactions, including hydrogen bonds with Trp_SDHB194 and Tyr_SDHD120, and a cation-π interaction with Arg_SDHC97. Leveraging this, we designed a SDH inhibitor E8 (enprocymid), exhibiting significant fungicidal activity (Ki = 0.019 μM) and reduced zebrafish toxicity (LC50 (96 h) = 1.01 mg a.i./L). This study elucidates the structure of fungal SDH and demonstrates the potential of ScSDH for rational design of next-generation fungicides, addressing fungal resistance and environmental toxicity in agriculture.
Subject terms: Enzymes, Proteins, Cryoelectron microscopy
Fungal diseases threaten crops worldwide. By resolving the cryo-EM structure of fungal succinate dehydrogenase, researchers designed an inhibitor with broad antifungal activity, strong field performance, and reduced environmental toxicity.
Introduction
Succinate dehydrogenase (SDH, EC 1.3.5.1), also known as Complex II, is a crucial enzyme in the respiratory chain, typically consisting of four subunits1 . As a key mitochondrial enzyme, SDH catalyzes a pivotal step in cellular energy metabolism by bridging the tricarboxylic acid (TCA) cycle and oxidative phosphorylation (OXPHOS)2 . The critical role of SDH in cellular energy metabolism has rendered it as a prime drug target, with its dysregulation implicated in a range of pathologies, including tumorigenesis and neurodegenerative disorders3–6 . Recent structural studies on SDH from Sus scrofa7,8 , Gallus gallus9,10 , Homo sapiens11 , Mycobacterium smegmatis12,13 and Tetrahymena thermophila14,15 have provided insights into SDH function, laying a foundation for therapeutic strategies targeting these conditions.
Beyond its biomedical significance, SDH has garnered attention in agriculture, where SDH inhibitors have become a focal point for fungicide development16,17 . This represents a promising frontier following the triazole and strobilurin fungicide eras. Notably, pydiflumetofen (PYD), a pyrazole amide fungicide developed by Syngenta in 2017, has gained traction in global agriculture since its launch18 . While PYD demonstrates broad-spectrum efficacy and has gained global agricultural adoption, its application is overshadowed by substantial ecological concerns, particularly pronounced toxicity to aquatic ecosystems. Studies report alarming acute toxicity to zebrafish, with a LC50 (96 h) of 0.729 mg a.i./L19 .
Structure-based drug design (SBDD) has become a cornerstone in the development of inhibitors for various targets20–23 . However, a critical gap persists: no high-resolution structure has been available for fungal SDH. This absence has severely hampered the rational SBDD of next-generation fungicides targeting this vital enzyme in pathogenic fungi, forcing reliance on non-fungal (e.g., porcine) SDH structures as proxies despite significant sequence and potential structural divergence in the fungicide-targeted membrane domains. Saccharomyces cerevisiae, sharing high homology with pathogenic fungi, serves as an ideal model to bridge this structural knowledge gap.
Here, we show the cryo-electron microscopy (cryo-EM) structures of SDH from Saccharomyces cerevisiae (ScSDH) in apo state and ubiquinone-1 bound states, revealing its type-D classification and heme-independent transfer mechanism. By determining the cryo-EM structure of the PYD-ScSDH complex at 3.23 Å resolution, we perform a structure-based rational design and identified a fungicide candidate E8 (The approved ISO common name is enprocymid), which demonstrates improved fungicidal efficacy and lower environmental toxicity. This study bridges the critical gap in fungal SDH structural biology and establishes a blueprint for developing eco-friendly fungicides with enhanced species selectivity.
Results and discussion
Purification and characterization of ScSDH
Native protein complexes isolated under physiologically relevant conditions are regarded as superior experimental specimens for high-resolution structural analysis. SDH is classified as a membrane protein, which presents significant challenges in its extraction and purification. To elucidate the three-dimensional architecture of ScSDH in its functional state, the enzyme was isolated from its native cellular environment through endogenous expression systems according to the method described by Graham24 . The mitochondrial membrane was collected by mechanical homogenization and differential centrifugation, and then further purified by ion exchange chromatography and gel filtration chromatography (Supplementary Fig. 1a). Blue native polyacrylamide gel electrophoresis (BN-PAGE, Supplementary Fig. 1b) showed that a single bond is present. SDS-PAGE (Supplementary Fig. 1c) and Liquid Chromatography Mass Spectrometer (LC-MS, Supplementary Table 1) successfully identified the four subunits of ScSDH, including SDHA, SDHB, SDHC, and SDHD. It confirmed that the purified protein was ScSDH. The activity of the purified ScSDH was further verified by the SDH activity assay method, and its Km was 195.13 ± 8.09 μM, kcat was 58.90 ± 0.36 min−1 (Supplementary Fig. 1d, Supplementary Table 2). The above results indicated that the SDH extracted from S. cerevisiae is a functional complex.
The overall architecture of ScSDH
Cryo-EM has been successfully utilized for structural analysis of membrane proteins in the fields of pesticides and pharmaceuticals, contributing to drug target discovery and mechanism research25–27 . Here, the ScSDH was determined by cryo-EM, with a total resolution of 3.36 Å (Fig. 1, Supplementary Fig. 2, Supplementary Table 3). ScSDH adopts a mushroom-like architecture, comprising a hydrophilic head (SDHA and SDHB subunits) and a hydrophobic membrane anchor (SDHC and SDHD subunits). All subunits assemble into a monomer, similar to porcine SDH7 .
Fig. 1. Overall architecture of the SDH from S. cerevisiae.
a Cryo-EM map of SDH at 3.36 Å resolution and cartoon representation of ScSDH. The SDH protein has a total length of 108 Å and a transmembrane portion of 40 Å. SDHA, SDHB, SDHC, and SDHD subunits are colored in blue, orange, green, and yellow individually. b Cartoon representation of SDHA subunits. FAD is shown as stick. All domains are color-coded. c Cartoon representation of SDHB subunits. FeS clusters are shown as spheres. All domains are color-coded. d Cartoon representation of hydrophobic transmembrane subunits SDHC and SDHD in front view and top view. The six transmembrane helices are labeled. PE phosphatidylethanolamine (shown as gray stick).
The SDHA subunit consists of four domains, flavin adenine dinucleotide (FAD)-binding domain, capping domain, helical domain, and C-terminal domain. The EM- density of FAD could be observed in the SDHA subunit, bound to the FAD-binding domain (Fig. 1b). The capping domain displays poorer density than the overall structure, likely due to its flexible loop-rich composition. Notably, the capping domain of ScSDH exhibits greater amino acid flexibility than other eukaryotic SDHAs, and this flexibility may arise from substrate or ligand binding at the FAD-capping domain interface28,29 . Adjacent to SDHA lies the butterfly-shaped SDHB subunit, housing three iron-sulfur clusters ([2Fe-2S], [4Fe-4S], [3Fe-4S]) across two domains (Fig. 1c). Like porcine SDHB, the [2Fe-2S] cluster is coordinated by an N-terminal loop and the others reside in the C-terminal domain, bound by conserved cysteines typical of SDHs. Overall, the soluble head (SDHA and SDHB) shares secondary structure features with known bacterial and mammalian SDHs7,11,30 .
The hydrophobic transmembrane anchor tail is situated opposite the hydrophilic head of SDH and contains two subunits, SDHC and SDHD (Fig. 1d). SDH diversity primarily stems from transmembrane regions, where subunit count (1, 2, 3), prosthetic groups (heme b, Fe-S cluster), and heme numbers vary (0, 1, 2)31 . In the realm of eukaryotic SDHs, the structural characteristics of membrane domain are broadly conserved. In ScSDH, both transmembrane subunits traverse the membrane three times. Unlike other eukaryotic SDHs, no clear heme density was observed between transmembrane subunits. This is consistent with the results of heme is nonessential for SDH function32,33 .
Phosphatidylethanolamine (PE) molecule was identified through density mapping, a finding that was corroborated by subsequent mass spectrometry analysis (Supplementary Fig. 1e). This PE unit was structurally located between the transmembrane helices II and III on the cytoplasmic face, with its tail extending into the membrane. In contrast, cardiolipin (CL), despite being detected by mass spectrometry, was conspicuously devoid in the map density profile. This interesting omission suggested that CL may be involved in highly dynamic structural interactions within protein complexes, thus avoiding traditional density mapping techniques. A parallel phenomenon also occurred in M. smegmatis SDH (PDB ID:7D6X)12 .
Heme is absent in ScSDH
The SDH superfamily comprises two enzymes—SDH and QFR (short for fumarate reductase)—that catalyze reversible metabolic reactions34 . These enzymes are characterized by their composition of four structurally analogous subunits. The SDH superfamily proteins are classified into six subtypes based on the number of transmembrane subunits and the heme content within the transmembrane domain. Previous work suggested that ScSDH is of type C35 . However, the structure of ScSDH resolved in this study revealed a lack of heme density, indicating that ScSDH was classified as type D, which differed from the SDH of most eukaryotes.
The cryo-EM structure of ScSDH provided a high-resolution density map of the transmembrane region, allowing for clear visualization and comparison of various amino acids and their side chains. A key structural difference between porcine SDH and ScSDH involves residue D108: tyrosine (Tyr_SDHD108) in ScSDH, which replaces histidine (His_SDHD102) in porcine SDH (Fig. 2). This substitution makes the distance between Tyr_SDHD108 and His_SDHC156 less than 4 Å, creating spatial hindrance that precludes heme binding. To verify the lack of heme in ScSDH, we measured the absorption spectra of ScSDH and porcine SDH. As shown in Fig. 3a, the reduced state of porcine SDH exhibits a characteristic absorption peak of reduced heme at 560 nm. No heme absorption peak was observed in either the oxidized or reduced state of ScSDH, but when heme was added, a heme absorption peak could be observed. To investigate the effect of heme deficiency on SDH activity, we detected ScSDH activity in the presence and absence of heme (Supplementary Fig. 1f), confirming that ScSDH maintains normal SDH activity. It is worth noting that in the presence of substrate ubiquinone-1 (UQ1), the kcat value of ScSDH significantly increased by about 10 times (Supplementary Table 2), while other species only increased by about 2 times, indicating that the ScSDH has excellent quinone mediated electron acceptance ability11,12 . The structural differences, spectral results, and activity detection collectively confirm that there is no heme in ScSDH and the absence of heme does not affect the activity of SDH enzyme.
Fig. 2. Comparison of transmembrane subunit structures between S. cerevisiae and S. scrofa.
a The alignment of ScSDH with S. scrofa SDH (Type C, PDB ID 1ZOY) in front view and top view. b Demonstration of heme-related amino acids in ScSDH with SsSDH. In SsSDH, D-102His and C-127His interact with heme, while in ScSDH, the allelic amino acids change to D-108Tyr and C-156His. c Top view of ScSDH and SsSDH. SDHC and SDHD alignment with surface representation. Sc, S. cerevisiae; Ss, S. scrofa. All subunits are color-coded. SDHC subunits are colored in green (Sc) and ice blue (Ss); SDHD subunits are colored in orange (Sc) and pink (Ss). PE and Heme are shown as gray stick.
Fig. 3. Structure of the UQ1 bound in S. cerevisiae SDH.
a Determination of heme in SDH by absorbance spectrometry. Reduced heme has a characteristic absorption peak at 560 nm. The presence of heme can be detected in the reduced SsSDH, while it cannot be detected in the reduced ScSDH. b The electron transfer pathway of ScSDH. Prosthetic groups in the path of electron transfer are labeled with edge-to-edge distances. Four prosthetic groups, FAD, [2Fe-2S], [4Fe-4S], and [3Fe-4S] are required for electron transfer flow from succinate to ubiquinone. c The interaction of UQ1 with ScSDH. The UQ1 is surrounded with B_190Pro, B_194Trp, B_239Ile, C_94Ser, C_97Arg and D_120Tyr.
The result that ScSDH lacks the heme contradicts long-standing biochemical assumptions33,36 . Early models were constructed based on sequence homology with heme-containing homologous proteins, assuming that Cys_SDHD109 forms a disulfide bond with porphyrin, while the adjacent Tyr_SDHD108 was modeled as an outward-facing conformation. Cryo-EM structural data revealed shows the opposite: Tyr_SDHD108 adopts an inward conformation, blocking the heme pocket the heme pocket (Fig. 2, Supplementary Fig. 3), which explains why heme is not bound in ScSDH.
Within the vast and varied expanse of the SDH superfamily proteins, the heme content within the transmembrane region has long been considered as a defining characteristic. However, recent findings have revealed that there are some rare exceptions to this rule. Two instances have been documented where the transmembrane segment lacks heme: the E. coli QFR (PDB ID: 1L0V)30 and the T. thermophila SDH (PDB ID: 8B6G)14 . The distinctiveness of E. coli QFR was marked by the absence of heme in its transmembrane region with an additional electron density, suggesting an alternative molecular occupancy that facilitates electron transfer between the quinone-binding sites Qp (a proximal quinone binding site) and Qd (a distal quinone binding site)30 . Moreover, cryo-electron tomography has revealed the structure of SDH in T. thermophila, where it is integrated as a component of a supercomplex15 . Despite the absence of heme in the transmembrane region, the soluble subunit SDHTT1 on the matrix side contained a bis-histidine C-type heme covalently bound by a single cysteine residue. This finding underscored the intricate relationship between protein structure and heme binding within the SDH superfamily.
Electron transfer pathway in ScSDH
All the prosthetic groups within the structure of ScSDH, including FAD, [2Fe-2S], [4Fe-4S], and [3Fe-4S] clusters, were unambiguously resolved in the cryo-EM density map (Fig. 3b, Supplementary Fig. 4g). The edge-to-edge distances between these redox-active prosthetic group, which were less than 14 Å, fell within a range conducive to efficient electron transfer. FAD was covalently anchored to the SDHA subunit, assuming the pivotal role of the initial electron acceptor within the complex. The catalytic activity of the complex was mediated by FAD, facilitating the oxidation of succinate and the concomitant reduction of fumarate37 . To gain a more intuitive understanding of electron transfer within SDH and to elucidate the reduction process of quinone, we determined a structural analysis of the complex formed by ScSDH and UQ1.
The quinone binding site was constructed by the highly conserved B chain, D chains, and the more variable C chain, exhibiting structural similarity to the Qp sites of human and porcine (Fig. 3c). The UQ1, the final electronic receiver, through an intricate hydrogen-bonding network, principally involving Trp_SDHB194 and Tyr_SDHD120, thus catalyzing the electron transfer essential for the metabolic vigor of the cell. The short distance (6.6 Å) between UQ1 and the [3Fe-4S] cluster facilitates rapid electron transfer.
In E. coli, porcine and human systems, succinate oxidation liberated electrons that were efficiently channeled towards ubiquinone through a complex redox cascade, comprising FAD, [2Fe-2S], [4Fe-4S] and [3Fe-4S] clusters. The final step in this electron transfer cascade was the delivery to ubiquinone, a process accelerated by heme b, which acted as a central electron pool38,39 . The strategic formation of semi-quinone at this stage was a key enhancement, enhancing both the rate and efficiency of electron transfer, thereby underscoring the intricate orchestration of biological energy conversion40 . In the SDH structure of M. smegmatis, Rieske type [2Fe-2S] replaced heme b as the electron pool, rather than completely lacking this type of cofactor12 . The structural conservation and redox center alignment between ScSDH and SDHs from E. coli, M. smegmatis, porcine and human sources suggested a commonality in their electron transfer mechanisms. Yet, the absence of heme b in ScSDH, as indicated by structural features and spectral experimental data, pointed to a more direct electron transfer pathway. Structural analyses further reveal that the electron transfer distance from [3Fe-4S] cluster to UQ1 in ScSDH (6.6 Å) is shorter than in porcine, M. smegmatis SDH1, and E. coli systems, promoting enhanced efficiency. While identical to human SDH in distance, the UQ1 of ScSDH adopts a distinct conformation stabilized by hydrogen bonding with Trp_SDHB194, diverging from human UQ1 binding. This finding challenges the earlier notion that heme b is a ubiquitous component of SDHs, reinforcing the idea that it is not an essential structural element for quinone reduction—a conclusion further supported by the ScSDH structure.
We propose heme absence in ScSDH reflects adaptation to its fermentative lifestyle, which reduces oxygen demand compared to other yeasts. Consequently, ScSDH evolved enhanced hypoxia tolerance, eliminating the need for the antioxidant role of heme.
Structural insights into the Q-binding site for fungicide discovery
SDH, a ubiquitous enzyme in mammals, bacteria, plants, and fungi, serves as a critical target in fungicide development. The quinone-binding sites of SDH exhibited species-specific structural diversity due to the non-conservatism of SDHC and SDHD. In porcine SDH and E. coli QFR, two distinct quinone-binding sites, Qp and Qd, have been identified, located on opposing sides of the membrane-spanning region7,30 . In M. smegmatis SDH2 trimer, the classical Qp site was blocked due to the presence of SDHF, forming the specialized QD1 and QD2 sites13 . However, the Qp site is still recognized as the classical inhibitor binding site.
PYD, a pyrazole amide fungicide developed by Syngenta, gained global agricultural acceptance since 2017. Its structure, featuring a difluoromethyl group at the pyrazole 3-position and an N-methoxy group, enhanced its control spectrum and potency41,42 . To elucidate the mechanism of action of PYD, we analyzed the cryo-EM structure of the PYD-ScSDH complex at 3.23 Å resolution (Fig. 4, Supplementary Fig. 5). Our analysis revealed that PYD bound via hydrogen bonds with Trp_194 in SDHB and Tyr_120 in SDHD, and its N-methoxy group formed additional hydrogen bond with Ser_94 in the C chain, enabling it to displace UQ1. Notably, the pyrazole ring of PYD formed a cation-π interaction with Arg_97 in SDHC (Fig. 4a). The bridging moieties of PYD formed an angle about 90°, clearly mapped in the cryo-EM density (Supplementary Fig. 5g).
Fig. 4. The binding site of Qp inhibitor PYD analysis in S. cerevisiae SDH.
a The binding mode between inhibitor PYD and target ScSDH in two views. PYD forms hydrogen bonds with B_194Trp, C_94Ser and D_120Tyr, and interacts with C_97Arg to form cation-π interactions. PYD is shown as stick, colored in yellow. b Comparison of porcine Qp sites bound to different inhibitors (including 3AEB, 3AE9, and 3AEE) with nearby amino acids showed that C_61Trp had different side chain orientations and underwent flipping in the presence of different inhibitors. c Inhibitor-bound Qp pocket and the alignments of Qp pocket between ScSDH (9KQ3) and SsSDH (3AEB). Due to changes in surrounding amino acids, the Qp pocket space in SsSDH is larger than that in ScSDH. Sc, S. cerevisiae; Ss, S. scrofa. 3AEB, 3AE9, 3AEE, and 9KQ3 are colored in blue, green, yellow, and purple.
Further analysis of amino acids within 5 Å of PYD showed that the inhibitor binding pocket is primarily surrounded by conserved B chain residues (Pro_190, Ser_191, Trp_193, Trp_194, His_237, Ile_239) and D chain residues (Asp_119, Tyr_120), along with the more variable C chain (Supplementary Figs. 6, 7). The C chain, a key determinant of structural diversity, includes residues Leu_81, Trp_90, Ser_93, Ser_94, Arg_97, and Ile_98, which are critical for inhibitor binding. Sequence alignment revealed that the inhibitor binding site on the ScSDHC subunit is relatively conserved in pathogenic fungi (Supplementary Figs. 6, 7). In addition, structure superposition also revealed structure similarities of Qp pocket between ScSDH and the predicted fungal SDHs (Supplementary Fig. 8)43 . This suggests that the ScSDH would be a promising model for guiding fungicides within a certain range.
Previously, the SDH complex from porcine was selected as a primary structural template for SDHI development, with its inhibitor bound complexes being extensively characterized through crystallographic and computational analyses44–46 . Comparative analysis of inhibitor binding models in porcine SDH revealed the openings of active cavity caused by variations of inhibitor. In porcine SDH, the flexibility of Trp_61 in the C chain emerges as a key factor determining the dimensions of active cavity. Variations in inhibitor size induce distinct angular shifts of the Trp_61 indole ring, enabling cavity size adaptation through dynamic structural rearrangements (Fig. 4b). In contrast, the ScSDH employs Trp_90 for similar functional adjustments. Structural superposition analysis (Fig. 4c) reveals that the ScSDH Trp_90 is spatially aligned with porcine SDH Met_65, which is located inside the cavity. Therefore, the ScSDH Trp_90 is a residue restricted by the proximal amino acid and has limited rotational freedom. This steric restriction results in a constitutively smaller Qp cavity in ScSDH compared to porcine SDH, highlighting the importance of considering species-specific amino acid compositions and spatial constraints in inhibitor design. The elucidation of the fungal ScSDH structure has established a structural framework for designing antifungal SDHIs with enhanced target specificity and reduced off-target effects in eukaryotic organisms.
Molecular design of SDH fungicide
The PYD was selected as the starting point to design SDH inhibitor. The complex structure of PYD-ScSDH was subjected to the CORE_GEN mode in ACFIS 2.0 tool to obtain a list of core fragments of ligand (Supplementary Fig. 9)47 . PYD was spliced into nine fragments and the binding free energy (ΔG) of them with SDH, as well as ligand efficiency (LE), were calculated. LE was a valuable method to measuring the fragment contribution for ΔG (Supplementary Fig. 9). The LE of fragment 1 was measured at 0.42, which rose to 0.53 in fragment 2 following the addition of one methoxy-amide bond. This result indicated that the amide bond in PYD was very important to maintain the activity. Among fragment 1 to 9, the fragment 3 and 4 gave out the higher LE than others fragments. Based on our previous study, the cyclopropyl attached to nitrogen atom would be favorable for increasing the compound activity against porcine SDH48 . Here, the fragment 10 and 11 were built based on fragment 3 and 4, respectively. It should be noted that the LE in fragment 10 was the highest with 1.20. Consequently, fragment 10 was selected as the core fragment.
Then, the 3D structure of fragment 10-bound ScSDH was uploaded to CAND_GEN module for fragment growing. At last, ACFIS 2.0 provided a series of fragment 10-derived compounds as potential SDH ligands (Supplementary Fig. 10). Among them, ligand 3 exhibited a lower calculated ΔG than PYD which was inferred to be bioactive toward SDH (Supplementary Table 4). Our previous studies had pointed out that the pyrazole ring attached to fluorine atom in the scaffold of SDH inhibitor was benefit for activity against SDH48–50 . Consequently, a series of pyrazole carboxamide with N-cyclopropyl compounds were synthesized (Fig. 5a).
Fig. 5. Design strategy of PYD analogs and binding modes with S. cerevisiae SDH.
a Design protocol of the PYD analogs. b The binding mode of E8 with ScSDH. E8 forms hydrogen bonds with B_194Trp and D_120Tyr, and interacts with C_97Arg to form cation-π interactions. c Overlapping diagram of E8 (blue stick) and PYD (yellow stick) combined with ScSDH. Their combination mode is similar.
All target compounds were evaluated against ScSDH. As shown in Table 1, most compounds exhibited good activity against ScSDH at 1 μM concentration. Notably, compound E8 demonstrated markedly superior potency (IC50 = 0.030 μM), surpassing the reference compound PYD (IC50 = 0.040 μM) by nearly 1.3-fold. These promising results warrant further in-depth biological evaluation of E8.
Table 1.
The inhibition constant of E8 and PYD against ScSDH
| No. | R1 | aI%/IC50 (μM) |
|---|---|---|
| E1 | H | 0.380 ± 0.033 |
| E2 | 2-Cl | 0.234 ± 0.013 |
| E3 | 3-Cl | 0.520 ± 0.016 |
| E4 | 4-Cl | 0.043 ± 0.004 |
| E5 | 4-methylsulfonyl | 7.610% |
| E6 | 2-CH3 | 0.164 ± 0.011 |
| E7 | 4-CH3 | 0.232 ± 0.017 |
| E8 | 2,4-Cl2 | 0.030 ± 0.003 |
| E9 | 3,5-Cl2 | 0.463 ± 0.033 |
| E10 | 2,3-F2 | 0.781 ± 0.067 |
| PYD | 0.040 ± 0.001 |
aI% = inhibition rate tested on 1 μM concentration.
The inhibition constant (Ki) represents the concentration of free inhibitor at which 50% of the enzyme is bound, and its value is independent of both enzyme and substrate concentrations used in the assay. Consequently, we define the inter-species selectivity of a compound as the ratio of its Ki for porcine SDH and ScSDH. As depicted in Table 2, the Ki of E8 against ScSDH, as well as porcine SDH, was performed. The results exhibited that E8 displayed 0.019 μM and 0.281 μM against ScSDH and porcine SDH, respectively, which were better than PYD. Among ScSDH and porcine SDH, the species selectivity for E8 was 14.78, which decreased to 8.28 for PYD. All these results indicated that E8 showed higher species selectivity than PYD. To investigate the mechanism of action of E8, the cryo-EM structure of E8-bound ScSDH (3.06 Å) was determined (Supplementary Fig. 11). The overall binding mode E8 was similar to that of PYD, forming hydrogen bond with Tyr_SDHD120 and Trp_SDHB194, cation-π interaction with Arg_SDHC97 (Fig. 5b, c). As shown Table 3, the double bond in E8 made the electronic energy (ΔEele) of it with ScSDH was more favorable than PYD, resulting in the ΔGcal of E8 lower than PYD. All these results were consistent with the experiment results.
Table 2.
The inhibition constant of PYD and E8 against SDH
| No. | Yeast | Porcine | Species selectivity |
|---|---|---|---|
| E8 | 0.019 ± 0.001 μM | 0.281 ± 0.014 μM | 14.78 |
| PYD | 0.042 ± 0.001 μM | 0.348 ± 0.017 μM | 8.28 |
Table 3.
Binding free energy (kcal/mol) of the E8 and PYD with ScSDH
| No. | ΔEvdw | ΔEele | ΔEPB | ΔESA | ΔH | -TΔS | ΔG | Ligand efficiency |
|---|---|---|---|---|---|---|---|---|
| E8 | −41.93 | −8.52 | 18.21 | −5.47 | −37.72 | 10.93 | −26.80 | 0.99 |
| PYD | −43.72 | −1.10 | 13.91 | −5.40 | −36.31 | 10.84 | −25.47 | 0.98 |
van der Waals energy (ΔEvdw), electrostatic energy (ΔEele), polar solvation energy (ΔEPB), nonpolar solvation energy (ΔESA), enthalpy change (ΔH), and entropy change (−TΔS).
Fungicidal activity of E8
Compound E8, which demonstrated excellent enzyme inhibitory activity, was systematically tested for its fungicidal activity. In vitro tests against representative plant-pathogenic fungi revealed the broad-spectrum efficacy of E8, with inhibition rates exceeding 70% at a low dosage of 1.56 mg/L, matching or surpassing PYD (Table 4). Spore germination assays demonstrate that E8 significantly inhibits the germination of 4 representative pathogenic species from different subphyla, confirming that this inhibitor impairs pathogenicity by blocking conidial germination (Supplementary Table 5). In vivo greenhouse experiments further demonstrated the significant activity of E8 against wheat powdery mildew (WPM) and cucumber powdery mildew (CPM) with EC50 values for E8 at 0.0462 mg/L for WPM and 0.5431 mg/L for CPM (Supplementary Table 6). Notably, E8 exhibited exceptional efficacy against WPM, showing over 20-fold greater activity than PYD. At low dose (0.097 mg/L), PYD-treated leaves exhibited numerous symptoms, whereas E8-treated showed significantly fewer symptoms, highlighting the superior efficacy of E8 even at lower application rates (Supplementary Fig. 12). These 12 plant-pathogenic fungi species span 3 subphyla, which suggests that diverse fungi species are susceptible to the E8.
Table 4.
The inhibition rates (%) of E8 against 10 phytopathogen species
| No. | Ss | Fg | Fp | Fo | Uv | Aa | As | Pi | Pc | Et |
|---|---|---|---|---|---|---|---|---|---|---|
| E8 | 74.67 | 97.31 | 85.33 | 87.33 | 100.00 | 95.67 | 80.00 | 81.97 | 76.33 | 100.00 |
| PYD | 75.67 | 98.32 | 84.67 | 92.00 | 100.00 | 95.00 | 82.69 | 88.10 | 76.67 | 100.00 |
All Fungicidal spectrum test at 1.56 mg/L dosage. Sclerotinia sclerotiorum (Ss); Fusarium graminearum (Fg); Fusarium pseudograminearum (Fp); Fusarium oxysporum (Fo); Ustilaginoidea virens (Uv); Alternaria alternata (Aa); Alternaria solani (As); Phytophthora infestans (Pi); Phytophthora capsica (Pc); Exserohilum turcicum (Et).
Field trials served as an effective approach to evaluate the control efficacy of compound under conditions approximating actual agricultural production. These trials assessed E8 against WPM, and fusarium head blight (FHB), a significant soil-borne disease that has adversely affected wheat yields and quality in recent decades51 . For WPM, the control effects were 92.38% for E8 with 187.5 g a.i./ha dosage, which was similar to PYD (93.35% control effects) (Supplementary Table 7). For FHB, E8 showed higher efficacy (88.81% control effect) at 150 g a.i./ha than that of validamycin (48.42% control effect), compared to tebuconazole (84.45% control effect) and PYD (89.08% control effect) (Fig. 6a, Supplementary Table 8).
Fig. 6. Protective activities of E8.
a Control effect of E8 against FHB in field trial. Three test plots were used per treatment. CK is no treatment. b DON toxin content and toxin inhibition of E8 treatments. Wheat samples collected using the 5-point sampling method were milled into flour, and the DON toxin content in the flour was assayed using HPLC, with the inhibition rate of the toxin calculated. c Average thousand grain weight and yield increase rate of E8 treatment. Test the weight of 1000 randomly selected wheat ears with the 5-point sampling method. d Growth of wheat ears of E8 treatment. Wheat ears infected with FBH turn black, while uninfected ears remain bright yellow. FBH fusarium head blight, DON deoxynivalenol, EC emulsifiable concentrate, SC suspension concentrate, AS aqueous solutions, CK control group.
The management of FHB present a dual challenge: controlling the disease itself and reducing the levels of deoxynivalenol (DON) produced by Fusarium species52 . As a result, lowering DON levels had become a critical standard for assessing the success of FHB management strategies. Significantly, E8 reduced DON levels and enhanced wheat yield (Supplementary Table 9). In the absence of fungicides, FHB can produce DON levels up to 1559.5 μg/kg. After applied E8, it reduced DON to 165.3 μg/kg at 180 g a.i./ha dosage (89.40% DON inhibition) and to 217.5 μg/kg at 150 g a.i./ha (86.05% DON inhibition), comparable to PYD (Fig. 6b). Additionally, E8 increased wheat yield by 10.18%, surpassing PYD (9.49%). These results highlight the potential of E8 as an SDH inhibitor for further development, effectively controlling FHB, reducing DON content, and enhancing yield (Fig. 6c, d).
The aquatic toxicity of a compound critically determines its developmental potential. Regulatory frameworks classify compounds with a LC50 (96 h) ≤ 1.0 mg a.i./L as highly toxic to aquatic organisms, imposing stringent restrictions on their use53 . While the reference compound PYD exhibits high toxicity (LC50 (96 h) = 0.729 mg a.i./L), precluding its application in rice fields, compound E8 demonstrates significantly reduced toxicity. Comprehensive evaluation classifies E8 as moderately toxic to zebrafish (LC50 (96 h) = 1.01 mg a.i./L) (Supplementary Table 10), indicating enhanced environmental compatibility. The structural comparison between PYD and E8 revealed a carbon-carbon double bond in E8. Previous analysis showed that similar carbon-carbon double bonds also exist in other agrochemicals, such as silthiofam and imazalil, and their carbon-carbon double bonds tend to be converted into dihydroxy structures during metabolism54,55 . Therefore, we postulated that the double bond of E8 may generate dihydroxy-containing metabolites in vivo or in environmental systems. The ecological structure-activity relationships model (ECOSAR) predicted that the proposed dihydroxy metabolite would exhibit low acute toxicity to fish (LC50 (96 h) = 344 mg/L) (Supplementary Fig. 13). This result is consistent with the reduced aquatic toxicity profile of E8, indicating that the double bond serves as a key structural determinant for detoxification in aquatic organisms.
In summary, the compound E8 has shown remarkable efficacy in fungicidal activity, as evidenced by greenhouse pot experiments and field trials, while also markedly decreasing toxicity to aquatic organisms. This makes it a highly promising candidate for a SDH inhibitor. Its strong inhibitory effects on various plant-pathogenic fungi, broad-spectrum fungicidal activity, and impressive results in field trials, along with its minimal toxicity to aquatic life, indicated that E8 has the potential to be a significant asset in future crop disease management.
In this work, a comprehensive structural analysis of ScSDH is present. The absence of heme in ScSDH and the identification of its electron transfer pathway underscore the diversity within the SDH superfamily and emphasize the importance of species-specific structural variations in inhibitor design. The structural insights into the quinone-binding site of ScSDH, provided a blueprint for developing SDHIs. The identification of E8, demonstrated the utility of ScSDH as a template for rational fungicide design, and offered practical solutions to address fungal resistance and environmental toxicity in modern agriculture. By providing a detailed structural framework for the design of more effective and eco-friendly fungicides, this study holds significant promise for enhancing crop protection while minimizing environmental impact, thereby playing a crucial role in promoting sustainable agricultural development. Future research should focus on exploring the structural diversity of SDH in other fungal species and optimizing inhibitors for practical agricultural applications.
Methods
The research complies with all relevant ethical regulations. The fish acute toxicity test protocol was J2022004-Z and approved by the animal ethics committee of Jiangsu Academy of Agricultural Sciences.
Yeast strain and culture
Saccharomyces cerevisiae strain, red star, cultured in YPG media (1% yeast extract, 2% tryptone and 3% glycerol) at 30 °C to an OD600 of 1.5 and then harvested by centrifugation at 3000 ×ばつ g for 5 min. Cell pellets were frozen at −80 °C until use.
Isolation of mitochondria
The cell pellets were thawed and suspended in buffer A (50 mM Potassium phosphate, pH 7.4, 1 mM EDTA, 0.9% KCl and 0.5 mM PMSF). Cells were lysed by high-pressure crusher at 1350 bar and then centrifuged at 3000 ×ばつ g for 10 min to remove cell debris. The pH of supernatant was adjusted to 5.4 with 1 M glycine-HCl, pH 2.35 and then centrifuged at 12,000 ×ばつ g for 45 min to collect crude mitochondrial pellets. And then pellets were resuspended with buffer B (100 mM Potassium phosphate, pH 7.4, 1 mM EDTA, 0.5 mM PMSF and 0.25 M sucrose) and centrifuged at 12,000 ×ばつ g. The pellets were identified as mitochondria and stored in a −80 °C.
Protein purification and characterization
The purification of ScSDH was carried out according to the method proposed by Graham24 . The isolated mitochondrial pellets were thawed and suspended in buffer B with homogenized. Mitochondrial membrane proteins were extracted with 1.6% (w/v) of the detergent sodium cholate for 3 h at 4 °C. Ammonium sulfate gradient salting out was used to purify ScSDH. ScSDH was precipitated at 45% to 60% ammonium sulfate saturation. The precipitate was resolubilized in buffer C (100 mM Potassium phosphate, pH 7.4, 1 mM EDTA and 0.5 mM PMSF) and diluted with buffer D (20 mM sodium phosphate pH 7.4, 0.1% DDM). The sample were loaded onto a Source 15Q column (GE Healthcare). The ScSDH was eluted from Source 15Q in buffer D with 200 mM KCl. The fractions contain ScSDH were incubated with inhibitor and then loaded onto a Superdex 200 increase column (GE Healthcare), which had been equilibrated in ×ばつPBS (Gibco) with 0.05% DDM. Elution was performed at a flow rate of 0.5 mL min−1. The peak fraction was collected and concentrated to 6 mg/mL.
The purified ScSDH were characterized with 4–16% BN-PAGE gel (Invitrogen), 15% SDS-PAGE and LC-MS56–58 .
Characterization of enzyme activity
The activity of SDH was characterized using the 2,6-dichlorphenolindophenol (DCIP, Sigma) method11,48 . Ubiquinone-1 (UQ1, MCE) was selected as the final electron acceptor to assess succinate dehydrogenase activity. The reaction system consisted of 20 mM phosphate buffer, 15 μM DCIP, 200 μM UQ1 and 0.01 to 1.5 mM succinate. The enzyme activity assay was performed at 30 °C by using the Cytation 5 (Biotek).
For the enzyme activity comparison of ScSDH with or without heme. The reaction system consisted of 20 mM phosphate buffer, 30 μM DCIP, 20 mM succinate, 0.394 μM heme (Sigma) and 0.05 to 1.0 mM UQ1.
Mass spectrometry
To identified the four subunits of ScSDH, following SDS-PAGE separation, 5-mm wide strip gel lanes were excised. The gel strips were washed in ddH2O and discolored using 50% acetonitrile (Sigma). Then, disulfide bonds were reduced using 20 mM dithiothreitol (Sigma) for 1 h at 55 °C, and subsequently free sulfhydryl groups were carbamidomethylated using 20 mM iodoacetamide (Sigma) for 30 min at room temperature in the dark. Trypsin (Promega) was added to digest samples at 37 °C for 14 h. The peptide samples were extracted using 65% acetonitrile and 5% formic acid in ddH2O and then dried under vacuum. Finally, Samples cleaned up using Oasis HLB cartridges (Waters) and extracted with 80% acetonitrile and dried. For LC-MS/MS Analysis, the dried peptide samples were redissolved in 30 μL of 0.1% formic acid in ddH2O (Thermo Scientific) and transferred into injection vials. Using an Easy nLC 1200 Rapid Separation Liquid Chromatography system (Dionex/Thermo Scientific, Germany), 3 μL of each sample were trapped on a C18 PepMap100 precolumn (particle size 3 μm; Dionex/Thermo Scientific) and analytical column (C18; particle size 2 μm; Thermo Scientific, Germany) with an aqueous-organic gradient (eluent A: 0.1% (v/v) formic acid in ddH2O; eluent B: 0.1% (v/v) formic acid/80% (v/v) acetonitrile/19.9%(v/v) ddH2O; The gradient was set as follows: 0–3 min, 8–10% B; 3–45 min, 10–25% B; 45–54 min, 25–45% B; 54–56 min, 45–99% B; 56–60 min, 99%B. Eluting peptides were electrosprayed (2.1 kV; transfer capillary temperature 275 °C) in positive ion mode into an Orbitrap Exploris480 tandem mass spectrometer equipped with a Nanospray. Total acquisition time 60 min according to the gradient. FT full MS (m/z 350 to 1200; resolution 60000 FWHM) and ACG Target 300, RF Lens 50%, Mass width ± 10 ppm, Isolation Window 1.6 m/z, HCD Collision Energies 30, Exclusion duration 45 s. For protein identification, Peak lists were extracted from fragment ion spectra using the Proteome Discoverer (PD) 3.0 (Thermo Scientific). Data were searched with Sequest HT against the red star protein database, including the sequence of SDH subunits. For each dataset, Acetyl (protein N terminus) and oxidation (Met) were chosen as variable modifications, Carbamidomethylation on cysteine (+ 57.021 Da) was specified as static modification. fragment mass tolerance was set to ±10 ppm, and two missed cleavage was allowed. high confidence corresponding to a false discovery rate (FDR) ± 1%.
Co-purified lipids from ScSDH were extracted by chloroform/methanol (2:1, v/v) and dried. Then the sample was re-dissolved in 60% acetonitrile. Analysis of samples was conducted on a SCIEX ExionLC system coupled with a SCIEX ZenoTOF 7600 mass spectrometer. Lipid analysis was carried out using reverse-phase (RP)-LC–MS in both positive and negative ESI modes. The separation was performed on a Kinetex C18 column (2.6 μm particle size) maintained at 45 °C with a flow rate of 0.30 mL/min. HPLC buffer A consisted of acetonitrile, methanol, and water (1:1:1, v/v/v) containing 5 mM ammonium acetate, while buffer B consisted of 100% isopropanol with 5 mM ammonium acetate. The gradient was set as follows: 0–0.5 min, 20% B; 0.5–1.5 min, 20–40% B; 1.5–3 min, 40–60% B; 3–13 min, 60–98% B; 13–14.5 min, 98% B; 14.6–17 min, 20%B; 1.5–3 min, 40–60% B. A 2-μL resultant solution was injected for LC–MS/MS analysis. MS/MS analysis was performed using the following parameters: ion spray voltage, 5.5 kV (+) and 4.5 kV (−); curtain gas, 35 psi; gas 1 and gas 2, 55 psi; declustering potential, 80 V (+/−); and interface heater temperature, 550 °C. Collision energy was set to 45 ± 20 V in negative modes, with dynamic background subtraction applied. The experiments were conducted with a 150 ms accumulation time for TOF MS (m/z 200–1200) and a 25 ms accumulation time for TOF MS/MS (m/z 70–1200) using information-dependent acquisition mode. MS-DIAL (Version 5.1) was employed to filter and identify candidate lipids using exact mass, retention time, and MS/MS fragmentation patterns. The SCIEX OS workstation (Version 3.3) was applied to verify MS/MS information and quantify lipid peak areas.
Spectroscopy
The buffer solution consisted of 20 mM PBS (pH 7.4), while the enzyme (ScSDH and porcine SDH) was utilized at a concentration of 8 μM. To initiate the experiment, the buffer and enzyme were thoroughly mixed to ensure homogeneity, following the light absorption in the range of 400 nm to 600 nm was monitored, which the air-oxidized signal was obtain. Subsequently, the reductant dithionite was introduced, which the dithionite-reduced signal was obtain.
Electron microscopy sample preparation and imaging
4 μL of 6.0 mg/mL ScSDH were applied to glow-discharged 300 mesh ANTcryoTM R1.2/1.3 grids. The grids were blotted 5 s with force 3 at 8 °C and 100% humidity. The images were collected on a Titan Krios G4 microscope operated at a voltage of 300 kV with a Falcon4 direct electron detector. Single-particle data acquisition was conducted with EPU package, with a calibrated magnification of ×ばつ165,000 which yields a final pixel size of 0.74 or 0.75 Å. A total dose on the detector was approximately 48.94, 49.87, 50.11, and 40.48 electrons per Å2 for ScSDH, ScSDH-UQ1 bound, ScSDH-PYD bound, and ScSDH-E8 bound. Each micrograph stack contains 32 frames.
Imaging processing
The images were processed using software CryoSPARC59 . The patch-based motion correction and CTF estimation were conducted using the Patch CTF algorithm. The particles were automatically picked and extracted with a box size of 384 pixels in CryoSPARC. The extracted particles were classified and screened with several 2D classification. Then Ab-initio reconstruction and heterogeneous refinement were performed with all selected particles. Well-sorted particles were finally subjected to non-uniform refinements, resulting in the generation of final maps. The resolution was estimated using the Fourier shell correlation (FSC) 0.143 criterion. All dataset processing can be found in Supplementary Figs. 2, 4, 5, 11 and Supplementary Table 3.
Model building and refinement
The ScSDH atomic model was constructed manually built in Coot60 using the predicted structures with AlphaFold. The real-space refinement of the model was carried out using Phenix with Ramachandran restraint61 and further manual adjustment was performed in Coot. The stereochemistry and geometry of the structure were evaluated using Molprobity62 . All reported resolution was estimated using the Fourier shell correlation (FSC) 0.143 criterion.
All figures were generated using the program UCSF Chimera63 , ChimeraX64 , and PyMOL (http://pymol.org).
Molecular design of PYD analogs
The PYD-bound ScSDH was set as the starting point to design SDH inhibitor through pharmacophore-linked fragment virtual screening (PFVS) strategy65 . The structure-based optimization design of SDH inhibitor was performed with ACFIS 2.0 tool47 . The CORE_GEN mode was used to obtain a list of computational core fragment in PYD based on the structure of PYD-bound ScSDH. Fragment 10, which had the largest increment in LE (ligand efficiency), was selected as the core fragment. Then, the three-dimensional structure of fragment 10-bound ScSDH was upload, followed by the CAND-GEN mode to grow ligand. At last, the ACFIS 2.0 provided a series pyrazole carboxamide with N-cyclopropyl compounds as potential SDH inhibitor. Among them, compound 3 exhibited a lower calculated binding free energy than PYD, indicating high potency against SDH. Based on our previous study, a series modified compound 3 were synthesized and preformed hit-to-lead optimization and eventually obtain compound E8.
Chemicals and instruments
All reagents and solvents were purchased from commercial suppliers and used without further purification. The reactions were performed under atmosphere. Reaction progress was generally monitored by thin-layer chromatography (TLC). Intermediates and target compounds were purified by silica gel column chromatography. 1H NMR, 13C NMR, and 19F NMR were obtained on a Varian Mercury-Plus 600 or 400 spectrometer (Varian Inc., Palo Alto, CA, USA), using chloroform (CDCl3), or dimethyl sulfoxide (DMSO) - d6 as deuterated solvents. The NMR data were analyzed using MestReNova 14.0.0 software. Melting points were measured on a BüCHI B-545 melting point apparatus without correction. High-resolution mass spectra (HRMS) was performed with an Agilent 6224 TOF LC/MS (Agilent Technologies, Santa Clara, CA, USA). The detailed synthetic method and characterization data of compounds E1-E10 (1H NMR, 13C NMR, 19F NMR, and HRMS) can be found in the Supplementary Figs. 14–53 and Supplementary Note 1.
Fungicidal activity in vitro
The in vitro fungicidal activity of compound E8 was evaluated in mycelia growth inhibition method. Sclerotinia sclerotiorum, Fusarium graminearum, Fusarium pseudograminearum, Fusarium oxysporum, Ustilaginoidea virens, Alternaria alternate, Alternaria solani, Phytophthora infestans, Phytophthora capsica, and Exserohilum turcicum were taken as the test objects. The commercial fungicides PYD was chosen as positive control. Fungal discs (5 mm diameter) were punched from the margins of colonies grown for 4 consecutive days. These discs were inoculated onto PDA plates containing a fungicide concentration of 1.56 mg/L, with three replicates per treatment. The plates were then incubated at 25 °C in darkness. The diameters of both the control and treated colonies were measured using the cross method, and the inhibition rate was calculated as Eq. 1:
Spore germination
To evaluate spore germination, fresh spore suspensions were evenly spread onto water-agar medium plates containing a series of gradient concentrations of E8 and PYD, with three replicates per concentration. After incubation at 28 °C for 12 h, spore germination rates at different concentrations were quantified. Germination was defined as the emergence of a germ tube reaching half the length of the minor of spore. The inhibition rate of spore germination was calculated using the Eq. 2:
The data were analyzed with DPS (Date Processing System, 9.5) and the EC50 values of the tested compounds were calculated.
Fungicidal activity in greenhouse
The fungicidal activities of E8 against wheat powdery mildew (WPM, Erysiphe graminis) and cucumber powdery mildew (CPM, Sphaerotheca fuliginea) were evaluated in greenhouse at 100 mg/L, 25 mg/L, 6.25 mg/L, 1.56 mg/L, 0.39 mg/L, and 0.097 mg/L. The commercial fungicides PYD was chosen as positive control. E8 and PYD were dissolved in DMF and diluted with distilled water containing 0.1% Tween 80 solution. Then the diluted solutions were sequentially gradient diluted to the test concentration. For each treatment, three potted wheat and cucumber plants of similar growth stages were selected. The solutions were uniformly sprayed on plant leaves. Following a 24-h drying period, S. fuliginea was inoculated onto cucumber leaves by the spore spray method, while E. graminis was inoculated onto wheat leaves by the spore shaking method. The plants were then placed in a suitable environment. The data were analyzed with Statistical Product and Service Solutions (SPSS) and the EC50 values of the tested compounds were calculated. The pot experiment was entrusted to Zhejiang A&F University for completion.
Fungicidal activity in field
The wheat field experiment against WPM (Erysiphe graminis, wheat variety: huaimai 33) and FHB (Fusarium graminearum, wheat variety: yannong 1212) was carried out in Huai’an (Jiangsu Province, China) in a plot with an area of 20 m2. The potential control efficacy was determined via the standard method. Compound E8 was prepared as 10% emulsifiable concentrate (EC) and diluted to 150, 187.5, and 225 g a.i./ha for WPM application, while it was diluted to 112.5, 150, and 180 g a.i./ha for FHB application. In the early onset of E. graminis, the wheat was uniformly sprayed with compound E8. For WPM control, PYD (20%, suspension concentrate, SC) and prothioconazole (30%, oil dispersion, OD) were used as a positive control, and water was used as a negative control. For FHB control, PYD (20%, SC), tebuconazole (430 g/L, SC), and validamycin (24%, aqueous solutions, AS) were used as a positive control, while water used as a negative control. Sprays were applied during early and peak anthesis of wheat, with efficacy against WPM and FHB assessed 14 and 17 days after the second spray, respectively. The disease index of WPM and FHB was determined by Eq. 3:
N = 9 when test WPM, N = 7 when test FHB.
The control efficacy was evaluated by Eq. 4:
The field trial was commissioned by the Plant Protection College of Nanjing Agricultural University for organization and implementation.
DON toxin determination methods: In each plot, samples were collected using the five-point sampling method. At each sampling point, 100 mature wheat spikes were harvested. The collected samples were mixed, rapidly air-dried, and threshed, then processed into flour. The deoxynivalenol (DON) content in the flour was quantitatively analyzed through liquid chromatography, with three replicates set for each plot. The inhibition rates were calculated as Eq. 5:
Average thousand grain weight determination methods: Samples were collected using the five-point sampling method in each plot. At each sampling point, 100 mature wheat spikes were collected. The collected wheat grains were mixed, rapidly air-dried, and threshed. Then 1000 wheat grains were randomly selected. 1000 wheat grains were weighed and each treatment was replicated three times. The yield increase rate was calculated as Eq. 6:
Acute toxicity zebrafish
The trial was conducted by Jiangsu Academy of Agricultural Sciences. To ensure animal welfare, all the animal maintenance and procedures were in accordance with recommendations established by the Animal Ethics Committee of Jiangsu Academy of Agricultural Sciences, as well as the guidance document of Code for Pesticide Registration Testing Quality Management and the Test Guidelines on Environmental Safety Assessment for Chemical Pesticides-Part 12: Fish Acute Toxicity Test (GB/T 31270.12-2014). The zebrafish (Brachydanio rerio H-B.) was juvenile (before reaching sexual maturity) and maintained at 22–23 °C on a 16–8 h light-dark cycle. Acute toxicity assessments of compound E8 on zebrafish were determined with a static test. For the definitive test, serial concentrations of 1.20, 1.00, 0.833, 0.694, 0.578, 0.482 mg a.i./L were set. Dechlorinated tap water (DTW) served as a blank control. Both experimental and control groups were set up with one replicate each, with each replicate consisting of 10 zebrafish. When an experiment resulted in varying concentrations with mortalities, the LC50 and the confidence limits (95%) were estimated using DPS.
Predicted toxicity of metabolite
The Ecological Structure Activity Relationships (ECOSAR) program package (Version 2.2) was used to predict the aquatic toxicity of E8 and its metabolite. The SMILES structure was imported into the program to estimate the acute toxicity to fish for chemicals of interest.
Reporting summary
Further information on research design is available in the Nature Portfolio Reporting Summary linked to this article.
Supplementary information
Source data
Acknowledgements
We are grateful for the financial support from the National Key Research and Development Program of China (2022YFD1700300, Q.W.), the National Natural Science Foundation of China (22477040, X.Z.). We thank Hai-Hong Zha from SCIEX, Analytical Instrument Trading Co., Ltd, Shanghai, China, for their kind help during the LC-MS data analysis. Ke-Dan Zhu from Jiangsu Academy of Agricultural Sciences, for their help during the fish acute toxicity test.
Author contributions
Y.Y. and G.Y. designed the experiments. Z.Li. purified the samples. Z.Li., Y.Y., collected and analyzed the cryo-EM data, built and refined the atomic models. Z.Li., X.Z., Y.Y. and G.Y. analyzed data and provided guidance and support. Y.W., J.W., E.F., G.C. and Y.F. performed activity assays and analyzed data. Y.W. performed the spectroscopy assays and analyzed data. X.Y. performed the protein and lipid identification and analyzed data. X.Z. and Z.Lu. carried out in molecular modeling. Y.H., G.W., and Q.W. synthesized and characterized the compounds used in this study. Z.Li., Y.H., X.Z. and Y.Y. wrote the manuscript. Q.W and X.Z. conceived the project and supervised the research.
Peer review
Peer review information
Nature Communications thanks the anonymous reviewers for their contribution to the peer review of this work. A peer review file is available.
Data availability
All the EM density map and structure reported in this study have been deposited in Protein Data Bank (PDB; http://www.rcsb.org) and the Electron Microscopy Data Bank (EMDB; https://www.ebi.ac.uk/pdbe/emdb), following the lists below: EMD-62490 and PDB 9KPS for the apo-ScSDH, EMD-62491 and PDB 9KPT for the ScSDH-UQ1, EMD-62495 and PDB 9KQ3 for the ScSDH-PYD, and EMD-63115 and PDB 9LIG for the ScSDH-E8. PDB codes of previously published structures used in this study are 1ZOY, 3AEE, 3AEB and 3AE9 of porcine; 7D6X of M. smegmatis Sdh1, 1L0V of E. coli, and 8B6G of T. thermophila. Source data are provided as a Source data file. Source data are provided with this paper.
Competing interests
The authors declare no competing interests.
Footnotes
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
These authors contributed equally: Zhi-Wen Li, Yuan-Hui Huang, Ge Wei.
Contributor Information
Xiao-Lei Zhu, Email: xlzhu@ccnu.edu.cn.
Ying Ye, Email: yeying@ccnu.edu.cn.
Guang-Fu Yang, Email: gfyang@ccnu.edu.cn.
Supplementary information
The online version contains supplementary material available at 10.1038/s41467-025-64001-0.
References
- 1.Bezawork-Geleta, A. et al. Mitochondrial complex II: at the crossroads. Trends Biochem. Sci.42, 312–325 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Vercellino, I. & Sazanov, L. A. The assembly, regulation and function of the mitochondrial respiratory chain. Nat. Rev. Mol. Cell Bio.23, 141–161 (2022). [DOI] [PubMed] [Google Scholar]
- 3.Aggarwal R. K. et al. Functional succinate dehydrogenase deficiency is a common adverse feature of clear cell renal cancer. Proc. Natl. Acad. Sci. USA 118, e2106947118 (2021). [DOI] [PMC free article] [PubMed]
- 4.Andrews, K. A. et al. Tumour risks and genotype-phenotype correlations associated with germline variants in succinate dehydrogenase subunit genes SDHB, SDHC and SDHD. J. Med. Genet.55, 384–394 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Farshbaf, M. J. & Kiani-Esfahani, A. Succinate dehydrogenase: prospect for neurodegenerative diseases. Mitochondrion42, 77–83 (2018). [DOI] [PubMed] [Google Scholar]
- 6.Eng, C., Kiuru, M., Fernandez, M. J. & Aaltonen, L. A. A role for mitochondrial enzymes in inherited neoplasia and beyond. Nat. Rev. Cancer3, 193–202 (2003). [DOI] [PubMed] [Google Scholar]
- 7.Sun, F. et al. Crystal structure of mitochondrial respiratory membrane protein complex II. Cell121, 1043–1057 (2005). [DOI] [PubMed] [Google Scholar]
- 8.Inaoka, D. K. et al. Structural insights into the molecular design of flutolanil derivatives targeted for fumarate respiration of parasite mitochondria. Int. J. Mol. Sci.16, 15287–15308 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Huang, L. S. et al. 3-nitropropionic acid is a suicide inhibitor of mitochondrial respiration that, upon oxidation by complex II, forms a covalent adduct with a catalytic base arginine in the active site of the enzyme. J. Biol. Chem.281, 5965–5972 (2006). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Huang, L. S., Lümmen, P. & Berry, E. A. Crystallographic investigation of the ubiquinone binding site of respiratory Complex II and its inhibitors. Biochim. Biophys. Acta (BBA)-Proteins Proteom.1869, 140679 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Du, Z. et al. Structure of the human respiratory complex II. Proc. Natl Acad. Sci. USA120, e2216713120 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Zhou, X. et al. Architecture of the mycobacterial succinate dehydrogenase with a membrane-embedded Rieske FeS cluster. Proc. Natl Acad. Sci. USA118, e2022308118 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Gong, H. et al. Cryo-EM structure of trimeric Mycobacterium smegmatis succinate dehydrogenase with a membrane-anchor SdhF. Nat. Commun.11, 4245 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Mühleip, A. et al. Structural basis of mitochondrial membrane bending by the I–II–III2–IV2 supercomplex. Nature615, 934–938 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Han, F. et al. Structures of Tetrahymena thermophila respiratory megacomplexes on the tubular mitochondrial cristae. Nat. Commun.14, 2542 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Luo, B. & Ning, Y. Comprehensive overview of carboxamide derivatives as succinate dehydrogenase inhibitors. J. Agric. Food Chem.70, 957–975 (2022). [DOI] [PubMed] [Google Scholar]
- 17.Zeng, L.-Q. et al. Comprehensive overview of the amide linker modification in the succinate dehydrogenase inhibitors. J. Agric. Food Chem.72, 26027–26039 (2024). [DOI] [PubMed] [Google Scholar]
- 18.Umetsu, N. & Shirai, Y. Development of novel pesticides in the 21st century. J. Pestic. Sci.45, 54–74 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Di, S. et al. Comprehensive evaluation of chiral pydiflumetofen from the perspective of reducing environmental risks. Sci. Total Environ.826, 154033 (2022). [DOI] [PubMed] [Google Scholar]
- 20.Andricopulo, A. D., Salum, L. B. & Abraham, D. J. Structure-based drug design strategies in medicinal chemistry. Curr. Top. Med. Chem.9, 771–790 (2009). [DOI] [PubMed] [Google Scholar]
- 21.Kitchen, D. B., Decornez, H., Furr, J. R. & Bajorath, J. Docking and scoring in virtual screening for drug discovery: methods and applications. Nat. Rev. Drug Discov.3, 935–949 (2004). [DOI] [PubMed] [Google Scholar]
- 22.Jin, Z. et al. Structure of Mpro from SARS-CoV-2 and discovery of its inhibitors. Nature582, 289–293 (2020). [DOI] [PubMed] [Google Scholar]
- 23.Forli, S. et al. Computational protein–ligand docking and virtual drug screening with the AutoDock suite. Nat. Protoc.11, 905–919 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Palmer, G. In Methods in Enzymology (Elsevier, 1978).
- 25.Chen, W. et al. Structural basis for directional chitin biosynthesis. Nature610, 402–408 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Yang, Z. et al. Identification, structure, and agonist design of an androgen membrane receptor. Cell188, 1589–1604.e1524 (2025). [DOI] [PubMed] [Google Scholar]
- 27.Lin, L. et al. Cryo-EM structures of ryanodine receptors and diamide insecticides reveal the mechanisms of selectivity and resistance. Nat. Commun.15, 9056 (2024). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Starbird, C. et al. New crystal forms of the integral membrane Escherichia coli quinol: fumarate reductase suggest that ligands control domain movement. J. Struct. Biol.202, 100–104 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Tomasiak, T. M., Maklashina, E., Cecchini, G. & Iverson, T. M. A threonine on the active site loop controls transition state formation in Escherichia coli respiratory complex II. J. Biol. Chem.283, 15460–15468 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Iverson, T. M. et al. Crystallographic studies of the Escherichia coli quinol-fumarate reductase with inhibitors bound to the quinol-binding site. J. Biol. Chem.277, 16124–16130 (2002). [DOI] [PubMed] [Google Scholar]
- 31.Lancaster, C. R. D. The di-heme family of respiratory complex II enzymes. Biochim. Biophys. Acta (BBA)-Bioenerg.1827, 679–687 (2013). [DOI] [PubMed] [Google Scholar]
- 32.Tran, Q. M. et al. Escherichia coli succinate dehydrogenase variant lacking the heme b. Proc. Natl Acad. Sci. USA104, 18007–18012 (2007). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Oyedotun, K. S., Sit, C. S. & Lemire, B. D. The Saccharomyces cerevisiae succinate dehydrogenase does not require heme for ubiquinone reduction. Biochim. Biophys. Acta (BBA)-Bioenerg.1767, 1436–1445 (2007). [DOI] [PubMed] [Google Scholar]
- 34.Jardim-Messeder, D. et al. Fumarate reductase superfamily: a diverse group of enzymes whose evolution is correlated to the establishment of different metabolic pathways. Mitochondrion34, 56–66 (2017). [DOI] [PubMed] [Google Scholar]
- 35.Lemire, B. D. & Oyedotun, K. S. The Saccharomyces cerevisiae mitochondrial succinate: ubiquinone oxidoreductase. Biochim. Biophys. Acta (BBA)-Bioenerg.1553, 102–116 (2002). [DOI] [PubMed] [Google Scholar]
- 36.Oyedotun, K. S., Yau, P. F. & Lemire, B. D. Identification of the heme axial ligands in the cytochrome b562 of the Saccharomyces cerevisiae succinate dehydrogenase. J. Biol. Chem.279, 9432–9439 (2004). [DOI] [PubMed] [Google Scholar]
- 37.Iverson, T. M. Catalytic mechanisms of complex II enzymes: a structural perspective. Biochim. Biophys. Acta (BBA)-Bioenerg.1827, 648–657 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Horsefield, R. et al. Structural and computational analysis of the quinone-binding site of complex II (succinate-ubiquinone oxidoreductase): a mechanism of electron transfer and proton conduction during ubiquinone reduction. J. Biol. Chem.281, 7309–7316 (2006). [DOI] [PubMed] [Google Scholar]
- 39.Tran, Q. M. et al. The quinone binding site in Escherichia coli succinate dehydrogenase is required for electron transfer to the heme b. J. Biol. Chem.281, 32310–32317 (2006). [DOI] [PubMed] [Google Scholar]
- 40.Anderson, R. F., Hille, R., Shinde, S. S. & Cecchini, G. Electron transfer within complex II: Succinate: ubiquinone oxidoreductase of Escherichia coli. J. Biol. Chem.280, 33331–33337 (2005). [DOI] [PubMed] [Google Scholar]
- 41.Stierli, D. et al. In Recent Highlights in the Discovery and Optimization of Crop Protection Products (Elsevier, 2021).
- 42.Wang, Z. et al. Evaluation of exploitive potential for higher bioactivity and lower residue risk enantiomer of chiral fungicide pydiflumetofen. Pest Manag. Sci.77, 3419–3426 (2021). [DOI] [PubMed] [Google Scholar]
- 43.Abramson, J. et al. Accurate structure prediction of biomolecular interactions with AlphaFold 3. Nature630, 493–500 (2024). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Ru, Y. et al. Discovery of novel nicotinamide derivatives by a two-step strategy of Azo-incorporating and bioisosteric replacement. J. Agric. Food Chem.72, 20794–20804 (2024). [DOI] [PubMed] [Google Scholar]
- 45.Li, H. et al. Design, synthesis, and fungicidal evaluation of novel N-methoxy pyrazole-4-carboxamides as potent succinate dehydrogenase inhibitors. J. Agric. Food Chem.71, 2610–2615 (2023). [DOI] [PubMed] [Google Scholar]
- 46.Wei, G. et al. Expanding the chemical space of succinate dehydrogenase inhibitors via the carbon–silicon switch strategy. J. Agric. Food Chem.69, 3965–3971 (2021). [DOI] [PubMed] [Google Scholar]
- 47.Hao, G.-F. et al. ACFIS: a web server for fragment-based drug discovery. Nucleic Acids Res.44, W550–W556 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Huang, Y.-H. et al. Structure-based discovery of new succinate dehydrogenase inhibitors via scaffold hopping strategy. J. Agric. Food Chem.71, 18292–18300 (2023). [DOI] [PubMed] [Google Scholar]
- 49.Huang, Y.-H. et al. Discovery of N-methoxy-(biphenyl-ethyl)-pyrazole-carboxamides as novel succinate dehydrogenase inhibitors. J. Agric. Food Chem.70, 14480–14487 (2022). [DOI] [PubMed] [Google Scholar]
- 50.Li, H., Wang, Y.-X., Zhu, X.-L. & Yang, G.-F. Discovery of a fungicide candidate targeting succinate dehydrogenase via computational substitution optimization. J. Agric. Food Chem.69, 13227–13234 (2021). [DOI] [PubMed] [Google Scholar]
- 51.Han, S. et al. Fusarium diversity associated with diseased cereals in China, with an updated phylogenomic assessment of the genus. Stud. Mycol.104, 87–148 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Wang, H. et al. Horizontal gene transfer of Fhb7 from fungus underlies Fusarium head blight resistance in wheat. Science368, eaba5435 (2020). [DOI] [PubMed] [Google Scholar]
- 53.Farah, I. F. et al. Pesticides in aquatic environment: occurrence, ecological implications and legal framework. J. Environ. Chem. Eng.12, 114072 (2024). [Google Scholar]
- 54.Authority, E. F. S. Peer review of the pesticide risk assessment of the active substance silthiofam. EFSA J.14, e04574 (2016). [Google Scholar]
- 55.Authority, E. F. S. Conclusion on the peer review of the pesticide risk assessment of the active substance imazalil. EFSA J.8, 1526 (2010). [Google Scholar]
- 56.Wittig, I., Braun, H.-P. & Schägger, H. Blue native PAGE. Nat. Protoc.1, 418–428 (2006). [DOI] [PubMed] [Google Scholar]
- 57.Hu, Q. et al. Oxidative lipidomics to elucidate the non-volatile derivatives of four typical triglycerides in vegetable oils under simulated frying conditions. Food Chem.410, 135414 (2023). [DOI] [PubMed] [Google Scholar]
- 58.Gong, H. et al. An electron transfer path connects subunits of a mycobacterial respiratory supercomplex. Science362, eaat8923 (2018). [DOI] [PubMed] [Google Scholar]
- 59.Punjani, A., Rubinstein, J. L., Fleet, D. J. & Brubaker, M. A. cryoSPARC: algorithms for rapid unsupervised cryo-EM structure determination. Nat. Methods14, 290–296 (2017). [DOI] [PubMed] [Google Scholar]
- 60.Emsley, P., Lohkamp, B., Scott, W. G. & Cowtan, K. Features and development of Coot. Acta Crystallogr., Sect. D: Biol. Crystallogr.66, 486–501 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Adams, P. D. et al. PHENIX: a comprehensive Python-based system for macromolecular structure solution. Acta Crystallogr., Sect. D: Biol. Crystallogr.66, 213–221 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Chen, V. B. et al. MolProbity: all-atom structure validation for macromolecular crystallography. Acta Crystallogr., Sect. D: Biol. Crystallogr.66, 12–21 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Pettersen, E. F. et al. UCSF Chimera—a visualization system for exploratory research and analysis. J. Comput. Chem.25, 1605–1612 (2004). [DOI] [PubMed] [Google Scholar]
- 64.Goddard, T. D. et al. UCSF ChimeraX: Meeting modern challenges in visualization and analysis. Protein Sci.27, 14–25 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Hao, G.-F. et al. Computational discovery of picomolar Q o site inhibitors of cytochrome bc 1 complex. J. Am. Chem. Soc.134, 11168–11176 (2012). [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
All the EM density map and structure reported in this study have been deposited in Protein Data Bank (PDB; http://www.rcsb.org) and the Electron Microscopy Data Bank (EMDB; https://www.ebi.ac.uk/pdbe/emdb), following the lists below: EMD-62490 and PDB 9KPS for the apo-ScSDH, EMD-62491 and PDB 9KPT for the ScSDH-UQ1, EMD-62495 and PDB 9KQ3 for the ScSDH-PYD, and EMD-63115 and PDB 9LIG for the ScSDH-E8. PDB codes of previously published structures used in this study are 1ZOY, 3AEE, 3AEB and 3AE9 of porcine; 7D6X of M. smegmatis Sdh1, 1L0V of E. coli, and 8B6G of T. thermophila. Source data are provided as a Source data file. Source data are provided with this paper.