Plant Pathol J > Volume 41(2); 2025 > Article
Huang, Jiang, Lin, Li, Bi, Qi, Tang, Wang, and Lin: Crystal Structure of an Aldo-keto Reductase MGG_00097 from Magnaporthe grisea

Abstract

The enzyme MGG_00097 from rice blast fungus (Magnaporthe grisea) is a NADPH-dependent oxidoreductase, involved in synthesizing glycerol from dihydroxyacetone phosphate and dihydroxyacetone. The 35.5-kDa monomer belongs to the aldo-keto reductase superfamily, characterized by a highly conserved catalytic tetrad. This study, elucidates the expression, purification, and kinetic properties of recombinant MGG_00097. The ternary complex of MGG_00097 with NADP+ and glycerol was refined to a 2.9 Å resolution, revealing critical insights into substrate binding and catalysis. NADP+ binds within the active site, with residues Ser221, Leu223, Ser225, Lys271, Ser272, Ser273, Thr274, Arg277, and Asn281 forming the substrate and cofactor-binding pockets. A Y56A mutation reveals the open conformation of the cofactor-binding pocket, with Glu29 and Gln226 functioning as hinge residues for the conformational changes upon cofactor binding. These findings contribute to the understanding of MGG_00097’s catalytic mechanism and offer a basis for further biochemical and potential biotechnological applications.

Rice (Oryza sativa) serves as a primary food source for more than half of the global population. However, rice blast disease, predominantly caused by the filamentous fungus (Magnaporthe grisea), leads to significant crop losses, with annual yield reductions ranging from 10-30% (Reddy et al., 2022; Zhang et al., 2018). The pathogen also affects over 50 other grass species, including cereals like wheat, rye, and barley (Fernandez and Orth, 2018; Pfeifer and Khang, 2018; Zhang et al., 2016, 2018). The infection process starts with spores adhesion, followed by the formation of a specialized structure called an appressorium (Skamnioti and Gurr, 2007). This structure accumulates glycerol, generating turgor pressure of up to 8 MPa, which is necessary for breaching the plant cuticle (Wang et al., 2003, 2005; Wilson and Talbot, 2009). Subsequently, fungal hyphae penetrate and colonize plant tissue, leading to cell death and crop destruction (Fernandez and Orth, 2018; Foster et al., 2017).
The genome of a rice pathogenic strain of M. grisea, 70-15, encodes a large and diverse set of secreted proteins (Dean et al., 2005). Analysis of the M. grisea genome indicates that germinating spores exhibit significant flexibility in their ability to produce glycerol in the appressorium from stored compounds (Dean et al., 2005). MGG_00097, originating from this M. grisea, has been identified as a specific NADPH-dependent oxidoreductase. The activities of NADH-dependent glycerol-3-phosphate dehydrogenase (GPD) and NADPH-dependent glycerol dehydrogenase (GD) have been reported in developing appressoria of M. grisea (Thines et al., 2000). MGG_00097 functions as a catalyst in the reversible reduction of various substrates, including ketones and aldehydes (Dean et al., 2005), which suggests it might play a crucial role in the process of producing glycerol. The rice blast fungus uses the pressure generated by glycerol as a physical force to break the rice leaves cuticle in order to infect the plant (De Jong et al., 1997).
MGG_00097 has been identified as a member of the aldo-keto reductase (AKR) family. AKRs are ubiquitous, comprising over 150 characterized enzymes that participate in diverse biological processes such as detoxification, osmotic regulation, and the metabolism of endogenous and xenobiotic compounds (Chen et al., 2023; Penning, 2015). Currently, about plant AKRs are being studied, with the most characterized being the AKR4C family involved in aldehyde detoxification and stress defense, osmolyte production, secondary metabolism, and membrane transport (Pan et al., 2019). For example, the AKR4C9 (Simpson et al., 2009) (At2g37770) from Arabidopsis thaliana (Arabidopsis) supports the importance of the enzymatic action directed towards oxidative stress-linked aldehydes, such as malondialdehyde. In addition, AKR4C7 from maize (Zea mays) catalyzes the oxidation of sorbitol to Glc (De Sousa et al., 2009).
To date, three-dimensional (3D) structures of AKRs have been determined in both apo and holo forms. All structural architectures share a common TIM barrel consisting of eight α-helices on the outside of the barrel and eight β-strands packed in the hydrophobic core of the enzyme (Jez and Penning, 2001). Here, we successfully determined the structures of MGG_00097 in its holo form bound to the cofactor NADP+, as well as in a ternary complex with both NADP+ and the substrate glycerol. The MGG_00097 displays a classical TIM barrel structure and a distinct NADP+ and glycerol binding conformation through crystal structure analysis. Phenylalanine residue (Phe55), serine residue (Ser172), tyrosine residue (Tyr220), and alanine residue (Ala254) are the critical determinants for substrate binding to the enzyme. This study presents the first crystallographic analysis of MGG_00097, providing a detailed structural and functional characterization.

Materials and Methods

Cloning, expression, and protein purification

The MGG_00097 wild-type (WT) and Y56A mutant genes were cloned into a pHis2 expression vector, which included an N-terminal His6-tag for purification purposes. The constructs were transformed into Escherichia coli BL21 (DE3) cells.
Cultures were induced with 0.2 mM IPTG (isopropyl-β-d-thiogalactopyranoside) at 16°C for 16 h. The bacterial culture was harvested through centrifugation at 2,000 ×g for 15 min at 4°C. The pellets were lysed in 150 mM NaCl, 25 mM Tris-HCl pH 8.0, 0.5 mM TCEP-HCl, and 1 mM PMSF by sonication. The lysate was then centrifuged at 25,000 ×g for 30 min at 4°C. Proteins were initially purified using Ni-NTA affinity chromatography and eluted with 25 mM Tris-HCl pH 8.0, 150 mM NaCl, 0.5 mM TCEP-HCl, and 250 mM imidazole pH 8.0. The eluate of target proteins was digested with TEV protease at 4°C overnight. TEV and His6 were removed by applying the solution to a Hi-Trap Q HP column. Peak fractions were pooled and further purified on a Superdex 200 10/300 GL column equilibrated with 25 mM Tris-HCl pH 8.0, 150 mM NaCl, and 0.5 mM TCEP-HCl. The final protein samples were flash-frozen in liquid nitrogen and stored at −80°C until further use.

Crystallization, data collection, and structure determination

Crystals of MGG_00097 (10 mg/ml) were obtained using the hanging-drop vapor diffusion method. The protein solution was mixed with well buffer containing 0.2 M NaCl, 2% ethyl acetate, 1.2 M NaH2PO4, and 1.2 M K2HPO4. The crystals were soaked in cryoprotectant and flash-cooled in liquid nitrogen for data collection at the Shanghai Synchrotron Radiation Facility (SSRF), which were subsequently processed using HKL2000 (Otwinowski and Minor, 1997) (https://www.hkl-xray.com/). X-ray diffraction data were processed using the PHENIX suite (Jumper et al., 2021; Varadi et al., 2022), and the structure was solved by molecular replacement using the Alphafold prediction model (AF-G4NEI2-F1) as the search model (Adams et al., 2010; Emsley and Cowtan, 2004). All structural representations were prepared with PyMOL software (https://www.pymol.org/). Comprehensive statistics for data collection, structure determination, and refinement are provided in Table 1. Coordinates and the structure factors have been deposited to Protein Data Bank (PDB) with accession codes 8YP9 (MGG_00097-NADP+) and 8YP2 (MGG_00097-Y56A-NADP+-Glycerol complex). The NCBI accession code for MGG_00097 is XP_003719043.

Site-directed mutagenesis and catalytic activity assay

Under standard assay conditions, the MGG_00097 (WT, 200 nM; F55A, 200 nM; S172A, 200 nM; Y220A, 200 nM; A254G, 200 nM) activity was spectrophotometrically measured in a reaction mixture containing 0-0.8 mM NADP+, 10 mM phosphate (pH 7.6), 150 mM NaCl, and 1% glycerol. The catalytic activity of the WT and mutant enzymes was assessed using a spectrophotometric assay to measure NADPH (abs = 340 nm) consumption. Kinetic parameters (Km and Kcat) were derived from Michaelis-Menten plots and analyzed using GraphPad Prism software (GraphPad Software Inc., San Diego, CA, USA).

Results and Discussion

Purification of aldo-keto reductase MGG_00097

Recombinant MGG_00097 was successfully purified to homogeneity, as indicated by a single band at approximately 35.5-kDa on sodium dodecyl sulfate-polyacrylamide gel electrophoresis (Fig. 1A). Size-exclusion chromatography confirmed the monomeric state of the protein in solution, consistent with the calculated molecular weight (Fig. 1A). The purified proteins obtained are deemed suitable for subsequent crystallization experiments.

Overall structures of MGG_00097

The crystal structure of the holo form of MGG_00097, complexed with NADP+, was resolved at 2.3 Å resolution. Only one molecule of MGG_00097 is present in the asymmetry unit. The protein adopts a classical TIM-barrel fold (also known as an (β/α)8 barrel) (Fig. 1B and C) (Brändén, 1991; Farber and Petsko, 1990; Komoto et al., 2004; Mindnich and Penning, 2009), a hallmark of the AKR family (Liu et al., 2014; Sengupta et al., 2015; Songsiriritthigul et al., 2020), which is mainly composed of eight α-helices (α1-α8) and eight β-strands (β1-β8) (Burton, 2018). In the MGG_00097-NADP+ structure, the central region is composed of eight β-strands, while the outer region is surrounded by eight α-helix (Fig. 1C). In addition to the core α-helices and β-strands, the structure of MGG_00097 also includes two secondary structure elements that are common to the AKR superfamily, namely the two auxiliary helices H1 (Pro240-Ser249) and H2 (Glu289-Gly301), pack outside the α-helical barrel (Fig. 1C).
Similar to other AKR proteins, the structure of MGG_00097 reveals three large exposed loops, loops A (between β5 and α4; His119-Asn149), B (between β8 and H1; Ser221-Asp239) and C (at the carboxyl-terminus; Arg302-Gln323) (Fig. 2), which are on the top of the barrel and are implicated in the determination of cofactor/substrate-binding site and the substrate specificity (Barski et al., 1996; Ma and Penning, 1999). The lengths and positions in sequence of these loops can vary among AKRs in other families.
The active site of aldo-keto reductases is conserved in both sequence and structure (Jez et al., 1997). Multiple sequence alignment shows that the active site of the enzyme preserves the conserved catalytic tetrad Asp51, Tyr56, Lys86, and His119 (Fig. 3). Interestingly, in all (β/α)8 barrel proteins, the general position of the active site is maintained, and this folding is thought to be the most common for enzymes (Farber and Petsko, 1990).

The crystal structure of MGG_00097Y56A-NADP+-glycerol

In order to understand more fully the structural characteristics of MGG_00097 that determine its substrate specificities, it would be desirable to acquire crystal structures of MGG_00097 bound with a substrate at its active site. In this present study, we initially endeavored to obtain the crystal of WT MGG_00097 binding with substrate, but was not successful. In MGG_00097, Asp51, Tyr56, Lys86, and His119 form the catalytic tetrad, where the tyrosine is the general acid-base involved in proton transfer (Schlegel et al., 1998). Thus, we attempted to obtain a trapped substrate or product/NADP+/enzyme complex using a Y56A mutant of MGG_00097. The Y56A mutation enabled the visualization of the open conformation of the cofactor-binding pocket. The binding of glycerol was confirmed within this pocket (Fig. 4A). Structural analysis indicated significant rearrangements in the side chains of Glu29 and Gln226 upon NADP+ binding, suggesting these residues act as hinges to facilitate the conformational changes necessary for cofactor binding (Fig. 4B). Aldo-keto reductases carry out a sequential ordered bi-bi catalytic mechanism in which the cofactor binds to the enzyme first and is released last in the reaction cycle (Penning, 2015; Sanli and Blaber, 2001). It has been proposed that the cofactor must bind to the AKR enzyme first in order to be buried at the bottom of the substrate binding chamber, and the substrate molecule then pile up on the nicotinamide ring of the NADP+ cofactor (Penning, 2015; Penning et al., 2019). The NADP+ molecule with anti-conformation is deep inside a channel connected to the central active site cavity in the ternary complex structure. The nicotinamide and ribose moieties are stabilized by potential hydrogen bonds interactions from the highly conserved residues Asp51, Ser172, and Asn173 (Fig. 4C). The phosphate group and the NADP+ adenine moieties are surrounded by Ser221, Lys271, Ser272, Ser273, Thr274, Arg277, and Asn281.
The crystal structures of MGG_00097-NADP+ (lightblue) and MGG_00097Y56A-NADP+-glycerol are superimposed (blue) (Fig. 5). Compared to the holo form, the Y56A mutation resulting in the opening of the substrate binding pocket. This conformational change allows for the pocket to be accessible to glycerol for binding (Fig. 5). While the overall conformation of MGG_00097Y56A is almost identical to the WT, notable conformational changes are observed in the side chains of Glu29 and Gln226. Upon NADP+ binding, the Gln226 side chain undergoes a rotation of ~60°, enabling it to tightly fit into the crescent groove of the extended NADP+ molecule. In contrast, the Glu29 side chain rotates approximately 30° in proximity to the glycerol molecule (Fig. 5).

Catalytic activity assay

The crystal structure of MGG_00097 were determined to investigate the effects of Ala mutations on the catalytic residues Phe55, Ser172, and Tyr220. Additionally, the effects of glycine mutations on Ala254 were also investigated in terms of catalysis (Fig. 6A). The apparent kinetic parameters of the Km-NADP+ and Km-glycerol pairs were measured and analyzed using Lineweaver-Burk plots (Dalziel, 1975; Jiang et al., 2016; Schuurink et al., 1990) (Fig. 6B-D). These biochemical analyses aimed to provide a deeper understanding of the changes in enzymatic activity caused by the alanine mutations from a biochemical perspective. The binding of NADP+ and glycerol to the enzyme significantly influences their respective Km values (Tables 2 and 3). In Table 3, the Km value of MGG_00097F55A for glycerol is significantly increased from 343.33 mM (WT) to 3,813 mM, indicating about elevenfold weaken binding affinity. Although Kcat slightly decreased by 1.38-fold compared to that of MGG_00097, Kcat/Km is decreased by 18.5-fold compared to that of the WT, showing that the catalytic efficiency of substrate glycerol reduction. The Km-NADP+ of MGG_00097F55A slightly increased from 0.07 mM (WT) to 0.08 mM. The Kcat-NADP+ of MGG_00097F55A decreased from 87.32/min to 7.51/min, and thus, catalytic efficiency (Kcat/Km) decreased by 12.5-fold. Based on the structure analysis, it is observed that Ser172 potentially interacts with NADP+ by hydrogen bonds. Upon mutation of Ser172 to alanine (MGG_00097S172A), the Km value for NADP+ increased from 0.07 mM to 0.14 mM. Additionally, the Kcat value for NADP+ decreased significantly from 87.32/min to 17.24/min. The results here shed the catalytic efficiency of the cofactor NADP+ is reduced in the MGG_00097S172A mutant.
In summary, we solved the crystal structures of MGG_00097 in cofactor bound forms (MGG_00097-NADP+ complex), as well as the MGG_00097Y56A-NADP+-glycerol ternary complex. The binary structure of MGG_00097-NADP+ shares the typical TIM fold and the NADP+ molecular with anti-conformation is deep inside a channel connected to the central active site cavity within the (β/α)8 barrel. Compared with previous studies of other AKRs, all other AKRs have the conserved catalytic tetrad residues, including AKR2E4 (D53, Y58, K87, and H120), AKR11C1 (D46, Y51, K76, and H119), and AKR11B4 (D54, Y59, K84, and H130) (Marquardt et al., 2005; Richter et al., 2010; Yamamoto and Wilson, 2013). In the MGG_00097 amino acid sequence, the conserved catalytic tetrad D51, Y56, K86, and H119 surrounded the substrate glycerol. The phosphate group forms a charged pair with D51. The histidine is believed to play a role in substrate orientation and transition state stabilization (Ehrensberger and Wilson, 2004; Khurana et al., 1998). A consensus among most AKRs is their preference for the coenzyme, NADPH over NADH, based on MGG_00097Y56A-NADP+ structural analysis, the residues K271, S272, S273, T274, and R277 contribute to a hydrogen bond network encompassing the 2’monophosphate of adenosine monophosphate (Abraham et al., 2022; Jez et al., 1997). These amino acids likely act as key contributing elements for MGG_00097’s NADP+ preference.
Structural comparison between MGG_00097-NADP+ and MGG_00097Y56A-NADP+-glycerol identified a conformational alteration of Glu29 and Gln226. The Y56A mutation causes the phenol group on tyrosine to be replaced by hydrogen ions, causes a conformational change in the protein and opens the substrate binding pocket to be accessible to glycerol. The F55A (amino acid lost π bond of benzene ring after mutation) and Y220A (amino acid changed from hydrophilic amino acid to hydrophobic amino acid after mutation) were located at both ends of glycerol molecule, and Km value was significantly increased after mutation, indicating that the binding affinity between protein and substrate glycerol was significantly weakened after mutation. These results indicate that F55 and Y220 are the key amino acid sites of substrate-binding pocket. Ser172 is the key residue in cofactor-binding pocket.
The appressoria of M. grisea generate enormous turgor pressure by accumulating glycerol at concentrations as high as 3M, enabling the appressorium to breach the plant cuticle. The synthesis of glycerol in eukaryotic cells can occur via several routes. In yeast, glycerol production from carbohydrates involves the enzyme GPD, which converts dihydroxyacetone phosphate to glycerol-3-phosphate in an NADH-dependent manner (Ansell et al., 1997). Glycerol can also be synthesized from dihydroxyacetone and glyceraldehyde by an NADPH-dependent reductases. In the filamentous fungus Aspergillus nidulans, these reactions are carried out by a single enzyme known as NADPH-dependent GD (Redkar et al., 1995). In the mycelium of M. grisea, the activity of GPD was low and did not show significant changes in response to hyperosmotic stress. However, the activities of dihydroxyacetone reductase and glyceraldehyde reductase were detected in the mycelium of M. grisea and increased approximately threefold under hyperosmotic stress (Thines et al., 2000). In developing appressoria, the activities of both enzymes were present and remained active even 48 h after the appressoria were fully mature (Thines et al., 2000).
In this study, we demonstrate that MGG_00097 exhibits NADPH-dependent oxidoreductase activity, specifically oxidizing glycerol to glyceraldehyde. NADPH-dependent oxidoreductases catalyze the reduction or oxidation of a substrate coupled to the oxidation or reduction, respectively. Thus, MGG_00097 may play a key role in supporting the normal life processes of M. grisea or in facilitating plant host infection as glyceraldehyde reductase. MGG_00097 crystal structures provided mechanistic details of cofactor-dependent AKR activity. AKRs encoding NADP(H)-dependent oxidoreductases exist in nearly all phyla and function in the phase 1 metabolism of endogenous substrates and xenobiotics (Jez et al., 1997). MGG_00097 might be a good candidate for the development of stress-tolerant rice and could play a vital role in agricultural technology.

Notes

Conflicts of Interest

No potential conflict of interest relevant to this article was reported.

Acknowledgments

This work was supported by National Natural Science Foundation of China Grants 32201025 (D.T.).

Fig. 1
The overall structure of MGG_00097. (A) Gel filtration (Superdex 200 10/300 GL) profile of MGG_00097. The horizontal axis is elution volume, and the vertical axis is ultraviolet (UV) absorption. The UV absorbance is shown as the blue line. The Coomassie blue stained sodium dodecyl sulfate-polyacrylamide gel electrophoresis gel shows the peak fraction of MGG_00097 from gel filtration. (B) The topology diagram of the MGG_00097 structure. β strands are indicated as arrows, and α helices are indicated as cylinders. β strands, α helices, and loop regions are colored red, blue, and light blue, respectively. (C) Stereo-view cartoon representation of MGG_00097 apo form showing the typical (β/α)8 fold. The light blue, MGG_00097. The N- and C-termini of MGG_00097 are labeled.
ppj-oa-07-2024-0115f1.jpg
Fig. 2
Molecular surface of the MGG_00097-NADP+ complex. Three large exposed loops are colored yellow. NADP+ is shown as green sticks. All figures were drawn using PyMOL.
ppj-oa-07-2024-0115f2.jpg
Fig. 3
Multiple sequence alignment of proteins in the aldo-keto reductase (AKR) family. Conserved catalytic tetrad DYKH are highlighted with red asterisks. AKR2E4: silkworm Bombyx mori AKR2E4A (UniProt KB: H9JTG9), AKR4C14: Oryza sativa L. ssp AKR4C14 (B8AC38), AKR11B4: Gluconobacter oxydans AKR11B4 (Q5FQJ0), AKR11C1: Bacillus halodurans AKR11C1 (P54569), AKR14A1: Escherichia coli AKR14A1 (Q46851). Conserved catalytic tetrad DYKH are highlighted with blue asterisks. The alignment was performed with ClustalW and ESPript.
ppj-oa-07-2024-0115f3.jpg
Fig. 4
MGG_00097 binding sites to NADP+ and glycerol. (A) The overall structure of MGG_00097-NADP+-glycerol complex (B) Interactions between glycerol and MGG_00097. (C) Interactions between NADP+ and MGG_00097. Glycerol is shown in pink stick and the interacting residues are shown as line; NADP+ is shown in green stick and the interacting residues are shown as line. Dashed lines show hydrogen bond network.
ppj-oa-07-2024-0115f4.jpg
Fig. 5
Structural superimposition of MGG_00097-NADP+ (colored in lightblue) and MGG_00097Y56A-NADP+-glycerol (colored in blue).
ppj-oa-07-2024-0115f5.jpg
Fig. 6
Structural observations of MGG_00097Y56A-NADP+-glycerol and reduction activity assays. (A) Structure demonstration of key residues F55, S172, Y220, and A254 binding to substrate. (B) Glycerol oxidation scheme. (C, D) The relative activity of MGG_00097 and its mutants saturation curve obtained by plotting the initial velocity of the reaction as a function of substrate concentration was analyzed using Michaelis-Menten equation, obtaining the kinetic parameters reported in Tables 2 and 3. Each unit of enzyme activity is defined as the amount of the enzyme that catalyzes the conversion of 1 mM of NADPH/min.
ppj-oa-07-2024-0115f6.jpg
Table 1
Statistics of crystallographic data collection and refinement
Dataset MGG_00097-NADP+ (PDB: 8YP9) MGG_00097-Y56A-NADP+-glycerol (PDB: 8YP2)
Space group P212121 P212121
Cell dimensions
 a, b, c (Å) 64.44, 112.24, 171.97 64.76, 110.78, 171.51
 α, β, γ (°) 90, 90, 90 90, 90, 90
 Wavelength (Å) 1.0000 1.0000
 Resolution (Å) 25.15-2.30 (2.33-2.30) 46.8-2.90 (2.95-2.90)
 Unique reflections 55,459 (5,272) 27,660 (1,297)
 Completeness (%) 99.3 (95.8) 98.8 (94.9)
 Redundancy 12.6 (10.6) 7.9 (6.5)
 Rpim (%) 8.2 (43.0) 14.5 (54.9)
 <I>/<σ(I)> 1.8 (1.2) 4.3 (1.0)
 CC1/2 0.988 (0.749) 0.921 (0.427)
Refinement
Rwork/Rfree (%) 17.83/22.74 (27.70/29.70) 20.25/24.85 (31.03/37.79)
Average B-factor (Å2)
 Protein 30.3 42.9
 NADP+ 35.3 47.5
 Glycerol - 55.6
R.m.s. deviation from ideality
 Bond length (Å) 1.403 1.207
 Bond angle (°) 0.012 0.006
Ramachandran plot
 Favored (%) 97.69 96.11
 Allowed (%) 2.31 3.68
 Outlier (%) 0 0.21

Values in parentheses are for the highest resolution shell. Rfree was calculated with 10% of the reflections selected randomly.

Table 2
Kinetic constants of MGG_00097WT, MGG_00097F55A, MGG_00097S172A, MGG_00097Y220A, and MGG_00097A254G with NADP+ as substrate
MGG_00097 (NADP+) WT F55A S172A Y220A A254G
Kcat (/min) 87.32 ± 7.46 7.51 ± 0.35 17.24 ± 1.12 22.65 ± 1.88 39.08 ± 4.04
Km (mM) 0.07 ± 0.007 0.08 ± 0.003 0.14 ± 0.02 0.07 ± 0.01 0.12 ± 0.003
Kcat/Km (/min/mM) 1208.76 ± 24.45 96.90 ± 3.79 126.70 ± 11.79 304.70 ± 14.41 316.88 ± 30.34
Table 3
Kinetic constants of MGG_00097WT, MGG_00097F55A, MGG_00097S172A, MGG_00097Y220A, and MGG_00097A254G with glycerol as substrate
MGG_00097 (glycerol) WT F55A S172A Y220A A254G
Kcat (/min) 255.43 ± 15.19 170.13 ± 24.01 74.44 ± 4.76 159.9 ± 4.58 145.3 ± 13.58
Km (mM) 343.33 ± 3.79 3813 ± 505.02 570.03 ± 6.04 801.43 ± 55.04 482.47 ± 8.54
Kcat/Km (/min/mM) 0.74 ± 0.047 0.04 ± 0.0006 0.13 ± 0.0073 0.20 ± 0.0082 0.30 ± 0.026

References

Abraham, N., Schroeter, K. L., Zhu, Y., Chan, J., Evans, N., Kimber, M. S., Carere, J., Zhou, T. and Seah, S. Y. K. 2022. Structure-function characterization of an aldo-keto reductase involved in detoxification of the mycotoxin, deoxynivalenol. Sci. Rep. 12:14737.
crossref pmid pmc
Adams, P. D., Afonine, P. V., Bunkóczi, G., Chen, V. B., Davis, I. W., Echols, N., Headd, J. J., Hung, L.-W., Kapral, G. J., Grosse-Kunstleve, R. W., McCoy, A. J., Moriarty, N. W., Oeffner, R., Read, R. J., Richardson, D. C., Richardson, J. S., Terwilliger, T. C. and Zwart, P. H. 2010. PHENIX: a comprehensive Python-based system for macromolecular structure solution. Acta Crystallogr. D Biol. Crystallogr. 66(Pt 2):213-221.
crossref pmid pmc
Ansell, R., Granath, K., Hohmann, S., Thevelein, J. M. and Adler, L. 1997. The two isoenzymes for yeast NAD+-dependent glycerol 3-phosphate dehydrogenase encoded by GPD1 and GPD2 have distinct roles in osmoadaptation and redox regulation. EMBO J. 16:2179-2187.
crossref pmid pmc
Barski, O. A., Gabbay, K. H. and Bohren, K. M. 1996. The C-terminal loop of aldehyde reductase determines the substrate and inhibitor specificity. Biochemistry 35:14276-14280.
crossref pmid
Brändén, C.-I. 1991. The TIM barrel: the most frequently occurring folding motif in proteins. Curr. Opin. Struct. Biol. 1:978-983.
crossref
Burton, Z. F. 2018. α/β Proteins. In: Evolution since coding: cradles, halos, barrels, and wings, eds. by Z. F. Burton, pp. 33-43. Academic Press, London, UK.
crossref
Chen, R., Wang, S., Sun, Y., Li, H., Wan, S., Lin, F. and Xu, H. 2023. Comparison of glyphosate-degradation ability of aldo-keto reductase (AKR4) proteins in maize, soybean and rice. Int. J. Mol. Sci. 24:3421.
crossref pmid pmc
Dalziel, K. 1975. Kinetics and mechanism of nicotinamide-nucleotid-linked dehydrogenases. Enzymes 11:1-60.
Dean, R. A., Talbot, N. J., Ebbole, D. J., Farman, M. L., Mitchell, T. K., Orbach, M. J., Thon, M., Kulkarni, R., Xu, J.-R., Pan, H., Read, N. D., Lee, Y.-H., Carbone, I., Brown, D., Oh, Y. Y., Donofrio, N., Jeong, J. S., Soanes, D. M., Djonovic, S., Kolomiets, E., Rehmeyer, C., Li, W., Harding, M., Kim, S., Lebrun, M.-H., Bohnert, H., Coughlan, S., Butler, J., Calvo, S., Ma, L.-J., Nicol, R., Purcell, S., Nusbaum, C., Galagan, J. E. and Birren, B. W. 2005. The genome sequence of the rice blast fungus Magnaporthe grisea. Nature 434:980-986.
pmid
De Jong, J. C., McCormack, B. J., Smirnoff, N. and Talbot, N. J. 1997. Glycerol generates turgor in rice blast. Nature 389:244.
crossref pdf
De Sousa, S. M., Rosselli, L. K., Kiyota, E., da Silva, J. C., Souza, G. H. M. F.n, Peroni, L. A., Stach-Machado, D. R., Eberlin, M. N., Souza, A. P., Koch, K. E., Arruda, P., Torriani, I. L. and Yunes, J. A. 2009. Structural and kinetic characterization of a maize aldose reductase. Plant Physiol. Biochem. 47:98-104.
crossref pmid
Ehrensberger, A. H. and Wilson, D. K. 2004. Structural and catalytic diversity in the two family 11 aldo-keto reductases. J. Mol. Biol. 337:661-673.
crossref pmid
Emsley, P. and Cowtan, K. 2004. Coot: model-building tools for molecular graphics. Acta Crystallogr. D Biol. Crystallogr. 60:2126-2132.
crossref pmid
Farber, G. K. and Petsko, G. A. 1990. The evolution of α/β barrel enzymes. Trends Biochem. Sci. 15:228-234.
crossref pmid
Fernandez, J. and Orth, K. 2018. Rise of a cereal killer: the biology of Magnaporthe oryzae biotrophic growth. Trends Microbiol. 26:582-597.
crossref pmid pmc
Foster, A. J., Ryder, L. S., Kershaw, M. J. and Talbot, N. J. 2017. The role of glycerol in the pathogenic lifestyle of the rice blast fungus Magnaporthe oryzae. Environ. Microbiol. 19:1008-1016.
crossref pmid pdf
Jez, J. M., Bennett, M. J., Schlegel, B. P., Lewis, M. and Penning, T. M. 1997. Comparative anatomy of the aldo-keto reductase superfamily. Biochem. J. 326:625-636.
crossref pmid pmc pdf
Jez, J. M. and Penning, T. M. 2001. The aldo-keto reductase (AKR) superfamily: an update. Chem. Biol. Interact. 130-132:499-525.
crossref pmid
Jiang, W., Wang, S., Wang, Y. and Fang, B. 2016. Key enzymes catalyzing glycerol to 1,3-propanediol. Biotechnol. Biofuels 9:57.
crossref pmid pmc
Jumper, J., Evans, R., Pritzel, A., Green, T., Figurnov, M., Ronneberger, O., Tunyasuvunakool, K., Bates, R., Žídek, A., Potapenko, A., Bridgland, A., Meyer, C., Kohl, S. A. A., Ballard, A. J., Cowie, A., Romera-Paredes, B., Nikolov, S., Jain, R., Adler, J., Back, T., Petersen, S., Reiman, D., Clancy, E., Zielinski, M., Steinegger, M., Pacholska, M., Berghammer, T., Bodenstein, S., Silver, D., Vinyals, O., Senior, A. W., Kavukcuoglu, K., Kohli, P. and Hassabis, D. 2021. Highly accurate protein structure prediction with AlphaFold. Nature 596:583-589.
crossref pmid pmc pdf
Khurana, S., Powers, D. B., Anderson, S. and Blaber, M. 1998. Crystal structure of 2,5-diketo-D-gluconic acid reductase A complexed with NADPH at 2.1-A resolution. Proc. Natl. Acad. Sci. U. S. A. 5:6768-6773.
Komoto, J., Yamada, T., Watanabe, K. and Takusagawa, F. 2004. Crystal structure of human prostaglandin F synthase (AKR1C3). Biochemistry 43:2188-2198.
crossref pmid
Liu, X., Wang, C., Zhang, L., Yao, Z., Cui, D., Wu, L., Lin, J., Yuan, Y-RA and Wei, D. 2014. Structural and mutational studies on an aldo-keto reductase AKR5C3 from Gluconobacter oxydans. Protein Sci. 23:1540-1549.
pmid pmc
Ma, H. and Penning, T. M. 1999. Conversion of mammalian 3α-hydroxysteroid dehydrogenase to 20α-hydroxysteroid dehydrogenase using loop chimeras: changing specificity from androgens to progestins. Proc. Natl. Acad. Sci. U. S. A. 96:11161-11166.
crossref pmid pmc
Marquardt, T., Kostrewa, D., Balakrishnan, R., Gasperina, A., Kambach, C., Podjarny, A., Winkler, F. K., Balendiran, G. K. and Li, X.-D. 2005. High-resolution crystal structure of AKR11C1 from Bacillus halodurans: an NADPH-dependent 4-hydroxy-2,3-trans-nonenal reductase. J. Mol. Biol. 354:304-316.
crossref pmid
Mindnich, R. D. and Penning, T. M. 2009. Aldo-keto reductase (AKR) superfamily: genomics and annotation. Hum. Genomics 3:362-370.
crossref pmid pmc pdf
Otwinowski, Z. and Minor, W. 1997. Processing of X-ray diffraction data collected in oscillation mode. Methods Enzymol. 276:307-326.
pmid
Pan, L., Yu, Q., Han, H., Mao, L., Nyporko, A., Fan, L., Bai, L. and Powles, S. 2019. Aldo-keto reductase metabolizes glyphosate and confers glyphosate resistance in Echinochloa colona. Plant Physiol. 181:1519-1534.
crossref pmid pmc
Penning, T. M. 2015. The aldo-keto reductases (AKRs): overview. Chem. Biol. Interact. 234:236-246.
crossref pmid pmc
Penning, T. M., Wangtrakuldee, P. and Auchus, R. J. 2019. Structural and functional biology of aldo-keto reductase steroid-transforming enzymes. Endocr. Rev. 40:447-475.
crossref pmid pmc pdf
Pfeifer, M. A. and Khang, C. H. 2018. A nuclear contortionist: the mitotic migration of Magnaporthe oryzae nuclei during plant infection. Mycology 9:202-210.
crossref pmid pmc
Reddy, B., Mehta, S., Prakash, G., Sheoran, N. and Kumar, A. 2022. Structured framework and genome analysis of Magnaporthe grisea inciting pearl millet blast disease reveals versatile metabolic pathways, protein families, and virulence factors. J. Fungi (Basel) 8:614.
crossref pmid pmc
Redkar, R. J., Locy, R. D. and Singh, N. K. 1995. Biosynthetic pathways of glycerol accumulation under salt stress in Aspergillus nidulans. Exp. Mycol. 19:241-246.
crossref pmid
Richter, N., Breicha, K., Hummel, W. and Niefind, K. 2010. The three-dimensional structure of AKR11B4, a glycerol dehydrogenase from Gluconobacter oxydans, reveals a tryptophan residue as an accelerator of reaction turnover. J. Mol. Biol. 404:353-362.
crossref pmid
Sanli, G. and Blaber, M. 2001. Structural assembly of the active site in an aldo-keto reductase by NADPH cofactor. J. Mol. Biol. 309:1209-1218.
crossref pmid
Schlegel, B. P., Jez, J. M. and Penning, T. M. 1998. Mutagenesis of 3 alpha-hydroxysteroid dehydrogenase reveals a “push-pull” mechanism for proton transfer in aldo-keto reductases. Biochemistry 37:3538-3548.
crossref pmid
Schuurink, R., Busink, R., Hondmann, D. H. A., Witteveen, C. F. B. and Visser, J. 1990. Purification and properties of NADP(+)-dependent glycerol dehydrogenases from Aspergillus nidulans and A. niger. J. Gen. Microbiol. 136:1043-1050.
crossref pmid
Sengupta, D., Naik, D. and Reddy, A. R. 2015. Plant aldo-keto reductases (AKRs) as multi-tasking soldiers involved in diverse plant metabolic processes and stress defense: a structure-function update. J. Plant Physiol. 179:40-55.
crossref pmid
Simpson, P. J., Tantitadapitak, C., Reed, A. M., Mather, O. C., Bunce, C. M., White, S. A. and Ride, J. P. 2009. Characterization of two novel aldo-keto reductases from Arabidopsis: expression patterns, broad substrate specificity, and an open active-site structure suggest a role in toxicant metabolism following stress. J. Mol. Biol. 392:465-480.
crossref pmid
Skamnioti, P. and Gurr, S. J. 2007. Magnaporthe grisea cutinase2 mediates appressorium differentiation and host penetration and is required for full virulence. Plant Cell 19:2674-2689.
crossref pmid pmc pdf
Songsiriritthigul, C., Narawongsanont, R., Tantitadapitak, C., Guan, H. H. and Chen, C. J. 2020. Structure-function study of AKR4C14, an aldo-keto reductase from Thai jasmine rice (Oryza sativa L. ssp. indica cv. KDML105). Acta Crystallogr. D Struct. Biol. 76:472-483.
crossref pmid
Thines, E., Weber, R. W. S. and Talbot, N. J. 2000. MAP kinase and protein kinase A-dependent mobilization of triacylglycerol and glycogen during appressorium turgor generation by Magnaporthe grisea. Plant Cell 12:1703-1718.
crossref pmid pmc
Varadi, M., Anyango, S., Deshpande, M., Nair, S., Natassia, C., Yordanova, G., Yuan, D., Stroe, O., Wood, G., Laydon, A., Žídek, A., Green, T., Tunyasuvunakool, K., Petersen, S., Jumper, J., Clancy, E., Green, R., Vora, A., Lutfi, M., Figurnov, M., Cowie, A., Hobbs, N., Kohli, P., Kleywegt, G., Birney, E., Hassabis, D. and Velankar, S. 2022. AlphaFold Protein Structure Database: massively expanding the structural coverage of protein-sequence space with high-accuracy models. Nucleic Acids Res. 50:D439-D444.
crossref pmid pmc pdf
Wang, Z.-Y., Jenkinson, J. M., Holcombe, L. J., Soanes, D. M., Veneault-Fourrey, C., Bhambra, G. K. and Talbot, N. J. 2005. The molecular biology of appressorium turgor generation by the rice blast fungus Magnaporthe grisea. Biochem. Soc. Trans. 33(Pt 2):384-388.
crossref pmid pdf
Wang, Z.-Y., Thornton, C. R., Kershaw, M. J., Debao, L. and Talbot, N. J. 2003. The glyoxylate cycle is required for temporal regulation of virulence by the plant pathogenic fungus Magnaporthe grisea. Mol. Microbiol. 47:1601-1612.
crossref pmid
Wilson, R. A. and Talbot, N. J. 2009. Under pressure: investigating the biology of plant infection by Magnaporthe oryzae. Nat. Rev. Microbiol. 7:185-195.
crossref pmid pdf
Yamamoto, K. and Wilson, D. K. 2013. Identification, characterization, and crystal structure of an aldo-keto reductase (AKR2E4) from the silkworm Bombyx mori. Arch. Biochem. Biophys. 538:156-163.
crossref pmid
Zhang, H., Zheng, X. and Zhang, Z. 2016. The Magnaporthe grisea species complex and plant pathogenesis. Mol. Plant Pathol. 17:796-804.
crossref pmid pmc
Zhang, S., Deng, Y. Z. and Zhang, L.-H. 2018. Phytohormones: the chemical language in Magnaporthe oryzae-rice pathosystem. Mycology 9:233-237.
crossref pmid pmc
TOOLS
METRICS Graph View
  • 0 Crossref
  •  0 Scopus
  • 279 View
  • 20 Download
ORCID iDs

Dan Tang
https://orcid.org/0009-0006-0986-9767

Related articles


ABOUT
BROWSE ARTICLES
EDITORIAL POLICY
FOR CONTRIBUTORS
Editorial Office
Rm,904 (New Bldg.) The Korean Science & Technology Center 22,
Teheran-ro 7-Gil, Gangnamgu, Seoul 06130, Korea
Tel: +82-2-557-9360    Fax: +82-2-557-9361    E-mail: paper@kspp.org                

Copyright © 2025 by Korean Society of Plant Pathology.

Developed in M2PI

Close layer
prev next