Articles Service
Research article
Engineering of the Phytase YiAPPA to Improve Thermostability and Activity and Its Application Potential in Dephytinization of Food Ingredients
Institute of Microbiology, Jiangxi Academy of Sciences, Nanchang 330096, Jiangxi Province, P.R. China
Correspondence to:J. Microbiol. Biotechnol. 2024; 34(8): 1660-1670
Published August 28, 2024 https://doi.org/10.4014/jmb.2403.03031
Copyright © The Korean Society for Microbiology and Biotechnology.
Abstract
Keywords
Graphical Abstract
Introduction
Phytate is the primary form of phosphorus in grasses and legumes [1, 2]. It can form stable and difficult-to-degrade complexes with proteins, minerals, and other substances in organisms, reducing the bioavailability of proteins and mineral elements [1, 2]. For this reason, phytate is considered an anti-nutritional factor. Phytase, also known as myo-inositol hexakisphosphate phosphohydrolase, can degrade phytate and generate inorganic phosphorus and phosphoinositides, eliminating the anti-nutritional effects of phytate [3, 4]. Owing to the lack of phytase in the human digestive system, phytate cannot be metabolized, which limits the absorption of nutrients in the intestine and reduces the nutritional value of food. Phytase can reduce the phytate level in food more effectively than cooking, boiling, fermentation, soaking, and other food processing methods. Adding phytase to plant-based food ingredients can degrade phytate and release inorganic phosphorus, which can improve the absorption and utilization of phosphorus in food, relieving the anti-nutritional effect of phytate [5-8]. Thus, phytase can be used as a new type of food additive in the food industry. However, numerous processing operations in the food industry necessitate execution under high temperatures. The limited thermostability of natural phytase poses a critical constraint in its practical utilization in food processing [9, 10]. Hence, the pursuit and advancement of improved thermostability of phytases remain a focal point of research among scholars worldwide.
Currently, protein engineering strategies used to enhance the thermostability of phytase include directed evolution (irrational design), semi-rational design, and rational design [11-14]. Rational design in the context of enzyme engineering involves directed modifications of enzymes based on the structure-activity relationship between molecular structure and function. Common methods employed in rational design include ion binding design, disulfide bond design, glycosylation site design, homology modeling, and molecular dynamics simulation [15-18]. For example, Sanchez-Romero
The strategy of protein surficial residue mutation was devised based on the above findings. Firstly, the GetArea software (http://curie.utmb.edu/getarea.html) was utilized to predict the localization of Gly and Lys residues on the protein surface. Next, each Gly residue was systematically substituted with Ala, and each Lys residue was replaced with Arg, leading to the construction of a mutant library. Directed mutagenesis was accomplished by carefully screening the resulting library to obtain mutants with specific modifications. Pang
The phytase YiAPPA (EC 3.1.3.2) derived from
Materials and Methods
Strains, Plasmid, and Culture Growth
Bioinformatic Analysis
The protein structure of the phytase YiAPPA was predicted using homology modeling by SWISS-MODEL (http://swissmodel.expasy.org) [24], with the molecular structure of the phytase YkAPPA from
Construction and Identification of Site-Specific Mutants
The primers were designed based on the base sequence of YiAPPA and the mutated amino acid sites by referring to the instructions of the QuickMutation kit, as shown in Table S1. Taking the construction of the site-directed mutant G16A as an example, the primers G16A-F and G16A-R were used to amplify linear fragments containing the vector sequence and the mutated gene sequence from the template recombinant plasmid pSTOP1622-
Induction, Expression, and Purification of Recombinant Phytase
The recombinant plasmid was transferred into competent
Determination of the Recombinant Phytase Activity
A 250-μl aliquot of the phytase solution (0.005 mg/ml) was diluted and thoroughly mixed with 750 μl of 0.25 mol/l sodium acetate buffer solution (pH 4.5). Thereafter, 2 ml of 1.5 mmol/l sodium phytate solution (0.25 mol/l sodium acetate buffer, pH 4.5) was added to the experimental group, and 2 ml of the end-point mixture (ammonium molybdate/ammonium vanadate/nitric acid) was added to the control group. The mixtures were thoroughly mixed and incubated at 37°C for 30 min. Subsequently, 2 ml of the end-point mixture was immediately added to the experimental group, and 2 ml of 1.5 mmol/l sodium phytate solution was added to the control group. All samples were thoroughly mixed and tested for their absorbance at 415 nm. The unit of phytase activity (U) was defined as the amount of phytase required to release 1 μmol/l of inorganic phosphorus from the 1.5 mmol/l sodium phytate solution per minute at pH 4.5 and 37°C.
Determination of the Enzymatic Properties of the Recombinant Phytase
The optimal reaction temperature, stability at 80°C, optimal reaction pH, stability at different pH levels, and protease resistance of the recombinant phytase were determined as previously reported [23].
The reaction system mentioned above was subjected to a 30-min reaction at temperatures ranging from 30°C to 90°C. The enzymatic activity of the samples was measured at various temperatures. The highest enzymatic activity was defined as 100%, and relative enzymatic activity was calculated for each temperature. A graph was plotted with relative enzymatic activity against temperature to determine the optimal reaction temperature. After incubating the enzyme solution at 80°C for 0–40 min, the enzymatic activity of the samples was determined according to the reaction system described above. The residual activity of the other samples was calculated using the enzymatic activity of the untreated sample as 100% and was plotted against temperature to evaluate the stability of the enzyme at 80°C.
The enzymatic activity at different pH levels (1.0–8.0) was determined. The relative enzymatic activity was calculated using the highest value as 100% enzymatic activity and was plotted against temperature to determine the optimal reaction pH. The enzyme solution was diluted using buffers of various pH levels (1.0–12.0), followed by a 2-h treatment at 37°C. Subsequently, the samples were diluted using the buffer with the optimal pH. The enzymatic activity of each sample was then measured based on the reaction system described above. The residual activity of other samples was calculated by using the enzymatic activity of the untreated sample as 100%, and the residual activity-temperature graph was plotted to evaluate the stability of enzymes under different pH conditions.
Solutions of 0.1 mg/ml of pepsin (0.25 mol/l Gly-HCl buffer, pH 2.0) and trypsin (0.25 mol/l Tris-HCl buffer, pH 7.0) were separately prepared. According to a proteinase-to-phytase mass ratio of 1:10, pepsin and trypsin were added to the phytase solution. The mixtures were then incubated at 37°C for 2 h, and a proteinase inhibitor was added to stop the reaction. Subsequently, the samples treated with proteinase were diluted 100-fold using the optimal pH buffer (0.25 mol/l sodium acetate buffer, pH 4.5). The enzymatic activity of the phytase before and after treatment with proteinase was determined using the reaction system described above.
Determination of Thermodynamic Parameters of the Recombinant Phytase
The thermodynamic parameter characterization of recombinant phytase mainly includes the determination of its half-life (
After incubating the enzyme solution (0.005 mg/ml) at 45°C, 50°C, 55°C, 60°C, 65°C, and 70°C for 30 min, the enzymatic activity of the samples was measured. Thereafter, the residual activity of the samples incubated at various temperatures was calculated with the enzymatic activity of the untreated sample as 100%. The temperature at which the residual activity was 50% was regarded as the
The
Determination of Kinetic Constant of the Recombinant Phytase
Sodium phytate solutions of various concentrations (0.0625%, 0.1%, 0.125%, 0.2%, 0.25%, 0.5%, 1.0%, and 1.5%) were prepared by diluting the stock solution with 0.25 mol/L sodium acetate buffer solution (pH 4.5). Equal volumes of phytase solution were added to each sodium phytate solution, and the enzymatic activity was determined following the method described above. Based on the double reciprocal plot, a graph was constructed with the reciprocal of substrate concentration as the x-axis and the reciprocal of enzyme activity as the y-axis. The slope of the line was
Molecular Dynamics Simulation and Molecular Docking
Molecular dynamics simulations were performed using Gromacs 2019.6 [32] at a temperature of 353 K (80°C) for a duration of 100 ns to analyze the thermodynamic fluctuations of the recombinant phytase. The simulations were conducted under constant temperature and pressure using the Amber14SB all-atom force field and the TIP3P water model. Periodic boundary conditions were applied during the simulations. The RMSD file reflecting the structure rigidity of the phytase molecule and the RMSF file reflecting the fluctuation of amino acid residues in phytase were obtained after the operation.
The recombinant phytase and phytate substrate were subjected to molecular docking using the software AutoDock 4.2.6 [33]. The binding energy and different energy contributions of the recombinant phytase and phytate substrate were calculated using the MMPBSA method [34]. Additionally, the software PyMOL v0.99 was employed to analyze the structural variations of the recombinant phytase at crucial amino acid residues.
Dephytinization of Food Ingredients
A 50-ml aliquot of 0.25 mol/l sodium acetate buffer (pH 4.5) was added to 5 g of food ingredients (wheat, corn, and soybean flours). After thoroughly mixing the samples, 10 U of purified recombinant phytase was added to the mixture. The mixture was then incubated at 55°C and 100 rpm for 4 h. Aliquots of samples were collected at 0.5, 1, and 2 h, and tested for the content of inorganic phosphorus and phytate as previously reported [35].
Statistical Analysis
Each of the enzymatic property tests was conducted in triplicate. Statistical analysis of experimental data was carried out using SigmaPlot 12.5, and all data are expressed as mean ± standard deviation.
Results and Discussion
Construction and Screening of Single-Site Mutants
This study intended to improve the thermostability of the phytase YiAPPA using protein surficial residue mutation. According to this molecular engineering strategy, the Lys and Gly residues located on the surface of YiAPPA were first analyzed. The amino acid sequence of YiAPPA was submitted to SWISS-MODEL for homology modeling using the molecular structure of the phytase YkAPPA (PDB ID: 4ARV) as the template, obtaining the molecular structure model for YiAPPA. The molecular structure model of YiAPPA was then submitted to the online prediction software GetArea. According to the working principle of GetArea, amino acid residues with a surface exposure ratio value greater than 50% were considered to be located on the surface of the protein molecule. The surficial Lys and Gly residues of YiAPPA predicted by GetArea are displayed in Table 1.
-
Table 1 . Surficial Lys and Gly residues of YiAPPA predicted using GetArea.
Amino acid Surface exposure ratio (%) Amino acid Surface exposure ratio (%) Gly16 100.0 Lys171 97.2 Lys35 92.9 Lys189 73.0 Gly81 72.7 Gly195 85.7 Gly86 67.3 Lys196 69.8 Gly117 100.0 Lys207 77.5 Gly119 77.3 Lys212 100.0 Lys121 79.7 Lys216 74.6 Lys129 96.4 Lys378 97.6 Lys130 68.1 Lys382 86.9 Lys144 70.8 Gly395 71.2 Lys152 72.2 Gly400 75.7 Gly159 81.2 Lys412 74.0 Lys167 99.4
Based on the prediction results in Table 1, YiAPPA mutants by replacing surface Lys with Arg (K35R, K121R, K129R, K130R, K144R, K152R, K167R, K171R, K189R, K196R, K207R, K212R, K216R, K378R, K382R, and K412R) and by replacing surface Gly with Ala (G16A, G81A, G86A, G117A, G119A, G159A, G195A, G395A, and G400A) were constructed. The heterologous recombination and expression of YiAPPA and its single-site mutants were carried out in
-
Fig. 1. SDS-PAGE analysis and enzymatic activity and thermostability determination of the recombinant phytase YiAPPA and its single-site mutants.
(A) M. Protein marker; YiAPPA, G16A, K35R, G81A, G86A, G117A, G119A, K121R, K129R, K130R, K144R, K152R, G159A, K167R, K171R, K189R, G195A, K196R, K207R, K212R, K216R, K378R, K382R, G395A, G400A, K412R, and K189R/K216R are purified recombinant phytases.
Fig. 1B shows that the specific activity of recombinant YiAPPA was 3960.81 U/mg. The specific activity of the K216R mutant was 4352.96 U/mg, which was 1.10 times higher than that of recombinant YiAPPA. The specific activity of the K35R, K129R, K130R, and K144R mutants was significantly lower than that of YiAPPA. The specific activity of the other mutants was almost the same as that of recombinant YiAPPA. The residual activity of the mutants K189R and K216R after 20 min at 80°C was 43.00% and 46.00%, respectively, significantly higher than that of recombinant YiAPPA (33.99%). The above results demonstrated that the molecular engineering of the phytase YiAPPA through protein surficial residue mutation generated two mutants: K189R, with significantly improved thermostability, and K216R, with enhanced enzymatic activity and thermostability.
Construction and Enzymatic Property Analysis of K189R/K216R
Construction of K189R/K216R and the Preliminary Characterization of Its Enzymatic Properties
Mutation stacking is a method to improve enzymatic properties by combining beneficial mutations [36, 37]. Based on the previous study results, in this study, two mutants with significantly enhanced thermostability were stacked, constructing a combined mutant K189R/K216R. The combined mutant K189R/K216R was introduced into
-
Table 2 . Specific activity (37°C), kinetic parameters (37°C), thermodynamic parameters, and protease resistance of the recombinant phytase.
Recombinant phytase Specific activity (U/mg) Km (μM)kcat (s-1)kcat /Km (μM-1 s-1)55°C t 1/2 (min)80°C t 1/2 (min)T 50 30 (°C)Tm (°C)Pepsin resistance (%) Trypsin resistance (%) YiAPPA 3960.81 ± 105.34 126.24 ± 14.18 10375.27 ± 1256.73 82.19 30.09 ± 0.19 14.81 ± 0.37 55.12 ± 0.87 48.36 ± 0.53 84.17 ± 3.20 82.28 ± 4.18 K189R/K216R 4469.13 ± 154.19 111.59 ± 11.03 10363.29 ± 987.56 92.87 60.45 ± 0.63 23.35 ± 0.57 62.44 ± 0.76 53.18 ± 0.62 85.43 ± 4.09 83.97 ± 3.57 K189R 3979.56 ± 121.00 125.30 ± 13.09 10301.32 ± 1107.83 82.21 47.34 ± 0.38 17.86 ± 0.51 57.32 ± 0.64 49.93 ± 0.32 84.36 ± 2.98 83.01 ± 5.03 K216R 4352.96 ± 121.15 115.01 ± 9.86 10381.09 ± 1091.32 90.26 50.09 ± 0.52 18.90 ± 0.43 59.85 ± 0.73 50.79 ± 0.57 84.91 ± 3.25 83.30 ± 4.18
Determination of the Optimal Reaction Temperature and Thermodynamic Parameters for K189R/K216R
This study determined the optimal reaction temperature and thermodynamic parameters of the recombinant phytase YiAPPA and its mutant K189R/K216R, and the results are shown in Fig. 2 and Table 2. Fig. 2A shows that the optimal reaction temperature of the recombinant phytase YiAPPA and K189R/K216R mutant was 55°C. The relative activity of YiAPPA was similar to that of K189R/K216R within the range of 30°C–55°C. However, at 60°C–90°C, the relative activity of K189R/K216R was significantly higher than that of YiAPPA. The above results indicate that K189R/K216R retained higher relative activity within the higher temperature range, and this may be associated with its superior thermostability. Fig. 2B reveals that the residual activity of K189R/K216R was significantly higher than that of YiAPPA after incubation at 55°C or 80°C, indicating its significantly higher stability at 55°C and 80°C than YiAPPA.
-
Fig. 2. Effect of temperature and pH on the activity and stability of the recombinant phytase.
(A) The effect of temperature on the activity of the recombinant phytase. (B) The effect of temperature on the stability of the recombinant phytase. (C) The effect of pH on the activity of the recombinant phytase. (D) The effect of pH on the stability of the recombinant phytase.
Based on these investigations, the thermodynamic parameter characterization of the recombinant phytase was carried out by determining the
Determination of the Optimal Reaction pH and pH Stability of K189R/K216R
This study determined the optimal reaction pH and the pH stability of the recombinant phytase YiAPPA and its mutant K189R/K216R, and the results are shown in Fig. 2C and 2D. The optimal reaction pH of the recombinant phytase YiAPPA and K189R/K216R mutant was 4.5, and they yielded similar relative enzyme activities at pH 1.0–8.0 (Fig. 2C). Within the pH range of 1.0–12.0, the stability of the recombinant phytase YiAPPA and K189R/K216R mutant was also similar (Fig. 2D). Within the pH range of 1.0–3.0, both exhibited over 80% residual enzyme activity. Within the pH range of 4.0–10.0, both demonstrated over 90% relative enzyme activity. However, within the pH range of 11.0–12.0, both were unstable and rapidly inactivated.
Determination of the Kinetic Parameters of K189R/K216R
The specific activity,
Determination of Protease Resistance of K189R/K216R
The results of the protease resistance analysis of the K189R/K216R mutant are shown in Table 2. The recombinant phytase YiAPPA and its mutants K189R, K216R, and K189R/K216R showed comparable protease resistance. Following pepsin treatment, recombinant YiAPPA retained 84.17% residual enzymatic activity, while the mutants K189R, K216R, and K189R/K216R retained 84.36%, 84.91%, and 85.43% residual enzymatic activity, respectively. After trypsin treatment, recombinant YiAPPA retained 82.28% residual enzymatic activity, whereas the mutants K189R, K216R, and K189R/K216R retained 83.01%, 83.30%, and 83.97% residual enzymatic activity, respectively. Overall, the introduction of mutations (K189R and/or K216R) in YiAPPA did not significantly affect its resistance to proteases.
Exploration of the Molecular Mechanism of Increased Thermostability and Activity of K189R/K216R
Molecular dynamics simulation of the recombinant phytase. At 353 K, a 100-ns molecular dynamics simulation was performed on the tertiary structure of YiAPPA and K189R/K216R to analyze their thermodynamic fluctuations, and the overall structural rigidity of the enzyme molecule and the fluctuations of amino acid residues were analyzed. The results are shown in Fig. 3. Fig. 3A shows that the RMSD value of YiAPPA and K189R/K216R fluctuated less after 40 ns and reached equilibrium, with the mean RMSD value of YiAPPA being 0.18 Å and that of K189R/K216R being 0.17 Å, that is, the mean RMSD value of the mutant K189R/K216R was slightly lower than that of YiAPPA. The RMSD value of an enzyme depends on its overall structural rigidity and stability. The lower the RMSD value, the greater the structural rigidity of the enzyme molecule, showing that it is more stable. The K189R/K216R mutant has a smaller RMSD value than YiAPPA, indicating it has a more stable molecular structure.
-
Fig. 3. Molecular dynamics simulation results of the recombinant phytase.
(A) RMSD value analysis of the recombinant phytase YiAPPA and K189R/K216R at 353 K (80°C). (B) RMSF value analysis of the recombinant phytase YiAPPA and K189R/K216R at 353 K (80°C).
Fig. 3B presents the RMSF analysis results of YiAPPA and K189R/K216R. At Lys-189 and its nearby residues (serine (Ser)-190 to cysteine (Cys)-198, S190-C198), the RMSF value of K189R/K216R decreased compared with that of YiAPPA, showing that the surrounding amino acid residues became more stable due to the mutation of Lys-189 in YiAPPA. This finding means that this single-site mutation also improved the stability of amino acid residues at nearby sites. In the region adjacent to the mutation site of Lys-216 (K216) (threonine (Thr)-209 to Cys-214, T209-C214, valine (Val)-217 to leucine (Leu)-219, V217-L219), the RMSF value of K189R/K216R was significantly lower than that of YiAPPA. However, the RMSF of the K216 site was unchanged, indicating that although Lys mutation at position 216 in YiAPPA, as an essential mutation site, did not alter its RMSF, it improved the stability of amino acid residues in the nearby region.
These findings demonstrate that K189R/K216R had a slightly lower average RMSD value than YiAPPA, and the RMSF value of its mutation and nearby sites was significantly reduced. The molecular structure of K189R/K216R was more rigid, and some of its amino acid residues were more stable, leading to the higher thermostability of K189R/K216R than that of YiAPPA.
Molecular structure analysis of the recombinant phytase. Using the structure of the phytase YkAPPA (PDB ID 4ARV) as the template, the structure models of YiAPPA and K189R/K216R were constructed through homologous modeling, and the interaction forces between the amino acid residues of mutation sites and their surrounding sites before and after mutation were analyzed to explore the reasons for the increased rigidity of K189R/K216R. The results are shown in Fig. 4A and 4B. Lys-189 is located in the α-helix away from the catalytic pocket and has a long and flexible side chain. When Lys-189 was mutated to Arg, a hydrogen bond was formed between the newly added Lys at position 189 and the Ala at position 184. Based on the results of the RMSF analysis, it can be concluded that the mutation of Lys to Arg at position 189 in YiAPPA further stabilized both itself and the Ser at position 190 to Cys at position 198. This increase in hydrogen bonding between the newly added R189 (Arg-189) and A184 (Ala-184) enhanced the conformational stability of this protein region. As shown in Fig. 4A and 4B, Lys-216 is located in the loop structure near the catalytic pocket, and its mutation to Arg introduced a hydrogen bond between Arg-216 and Ser-218. The results of molecular dynamics simulation revealed significantly reduced RMSF values for Thr-209 to Cys-214 and Val-217 to Leu-219. Specifically, this decrease indicates enhanced rigidity in the loop structure (Thr-209 to Cys-214) and the β-sheet structure (Val-217 to Leu-219). Considering the RMSF result analysis, it is apparent that the newly formed hydrogen bond between R216 (Arg-216) and S218 (Ser-218) improved the stability of both the aforementioned loop structure and the β-sheet structure. Pang
-
Fig. 4. Molecular structure analysis of the recombinant phytase.
(A) Molecular structure analysis of the recombinant phytase YiAPPA. (B) Molecular structure analysis of the recombinant phytase K189R/K216R. The blue parts in the figure represent the amino acid residues constituting the active site of the phytase. The red parts indicate the mutated amino acid residues (Lys-189 mutated to Arg-189; Lys-216 mutated to Arg-216) and the amino acid residues forming new hydrogen bonds with them (Ala-184 forms a new hydrogen bond with Arg-189; Ser-218 forms a new hydrogen bond with Arg-216). Hydrogen bonds are outlined in the magenta dashed box.
To further explore the molecular mechanism underlying the increased enzymatic activity of the mutant, the MMPBSA method was employed to calculate and decompose the total binding free energy of recombinant phytase and substrate phytate. Solvation energy refers to the sum of polar and non-polar solvation energy. The solvation energy values of YiAPPA and its mutants K189R, K216R, and K189R/K216R were 2055.79, 2072.31, 2098.97, and 2296.54 kcal/mol, respectively (Table 3). Theoretically, higher solvation energy is not conducive to substrate and enzyme binding [38]. The solvation energy of K216R and K189R/K216R, which displayed significantly improved enzymatic activity, was greater than that of YiAPPA. However, the absolute values of van der Waals energy and electrostatic energy for the two mutants were higher than those of YiAPPA, which was sufficient to offset the adverse effects of solvation during enzyme-substrate binding. The total binding energy of the K189R mutant was -67.29 kcal/mol, similar to that of YiAPPA. The total binding energy of the K216R mutant was -92.07 kcal/mol, which decreased by 43.41% compared to that of YiAPPA. The total binding energy of the K189R/K216R mutant was -97.07 kcal/mol, which was reduced by 51.20% relative to YiAPPA. These results show that the mutants K216R and K189R/K216R have increased stability and affinity after substrate binding, which is consistent with the changing trend of
-
Table 3 . Binding energy and various energy contributions of recombinant phytase and substrate phytate (kcal/mol).
Recombinant phytase ΔGvdW ΔGele ΔGpolar ΔGnon-polar ΔGbind WT -18.35 ± 2.70 -2101.65 ± 13.67 2059.44 ± 9.03 -3.65 ± 0.04 -64.20 ± 16.60 K189R -21.96 ± 2.81 -2117.64 ± 7.96 2075.72 ± 2.09 -3.41 ± 0.08 -67.29 ± 8.69 K216R -18.59 ± 2.53 -2172.44 ± 8.28 2103.10 ± 2.77 -4.13 ± 0.03 -92.07 ± 9.09 K189R/K216R -19.51 ± 1.93 -2374.11 ± 8.52 2300.54 ± 15.18 -4.00 ± 0.02 -97.07 ± 17.51
-
Fig. 5. The nucleophilic attack distance between His28 in recombinant phytase and phytate substrate.
(A) YiAPPA. (B) K189R. (C) K216R. (D) K189R/K216R. The red dotted line indicates atomic distance.
Application of the Recombinant Phytase as a Biocatalyst in Dephytinization of Food Ingredients
The phytate in food ingredients (wheat, corn, and soybean flours) was hydrolyzed by the recombinant phytase at 55°C. After hydrolysis, the content of inorganic phosphorus in the sample and the phytate degradation rates were measured and shown in Table 4. After the treatment of recombinant phytase YiAPPA and its K189R/K216R mutant, the content of inorganic phosphorus in the food ingredients and its phytate degradation rate were enhanced. In addition, compared with that of YiAPPA, the phytate hydrolysis efficiency of K189R/K216R (based on the inorganic phosphorus content in samples) against the food ingredients was increased by 2.36 (wheat flour), 2.28 (corn flour), and 1.73 times (soybean flour), respectively. After treatment with K189R/K216R at 55°C for 2 h, the inorganic phosphorus content in the wheat, corn, and soybean flour samples was 20.14 (wheat flour), 13.04 (corn flour), and 27.54 mg/g (soybean flour) with phytate degradation rates of 79.12%, 57.59%, and 81.52%, respectively. The above results reveal that the K189R/K216R mutant could more effectively reduce phytate in food ingredients and release inorganic phosphorus than YiAPPA, eliminating the anti-nutritional effect of phytate and improving the absorption and utilization of phosphorus in food.
-
Table 4 . Hydrolysis of phytates in food ingredients by the recombinant phytase.
Food Ingredient Recombinant Phytase Inorganic phosphorus (mg/g) Phytate degradation rate (%) 0.5 h 1 h 2 h 0.5 h 1 h 2 h Wheat flour YiAPPA 5.09 ± 0.39 7.03 ± 0.28 8.52 ± 0.67 19.99 27.62 33.47 K189R/K216R 6.65 ± 0.53 15.05 ± 0.56 20.14 ± 0.98 26.13 59.13 79.12 Corn flour YiAPPA 3.49 ± 0.36 4.89 ± 0.42 5.72 ± 0.37 15.41 21.60 25.26 K189R/K216R 4.58 ± 0.39 9.92 ± 0.51 13.04 ± 0.49 20.23 43.81 57.59 Soybean flour YiAPPA 8.94 ± 0.38 14.08 ± 0.51 15.95 ± 0.63 26.46 41.68 47.21 K189R/K216R 10.08 ± 0.59 21.31 ± 0.74 27.54 ± 0.66 29.84 63.08 81.52
This study successfully created the phytase mutant K189R/K216R with improved thermostability through protein surficial residue mutation. The
Supplemental Materials
Acknowledgments
This work was financially supported by grants from the National Natural Science Foundation of China (grant number 32160579) and Jiangxi Natural Science Foundation (grant numbers 20212BCJ23033, S2021GDQN2403, and 2022YJC2006).
Conflict of Interest
The authors have no financial conflicts of interest to declare.
References
- Philippi H, Sommerfeld V, Windisch W, Olukosi OA, Monteiro A, Rodehutscord M. 2023. Interactions of zinc with phytate and phytase in the digestive tract of poultry and pigs: a review.
J. Sci. Food Agric. 103 : 7333-7342. - Handa V, Sharma D, Kaur A, Kumar Arya S. 2020. Biotechnological applications of microbial phytase and phytate in food and feed industries.
Biocatal. Agric. Biotechnol. 25 : 101600. - Hussain SM, Hanif S, Sharif A, Bashir F, Iqbal HMN. 2022. Unrevealing the sources and catalytic functions of phytase with multipurpose characteristics.
Catal. Lett. 152 : 1358-1371. - Rizwanuddin S, Kumar V, Naik B, Singh P, Mishra S, Rustagi S,
et al . 2023. Microbial phytase: their sources, production, and role in the enhancement of nutritional aspects of food and feed additives.J. Agric. Food Res. 12 : 100559. - Song HY, El Sheikha AF, Hu DM. 2019. The positive impacts of microbial phytase on its nutritional applications.
Trends Food Sci. Technol. 86 : 553-562. - Singh B, Kumar G, Kumar V, Singh D. 2021. Enhanced phytase production by
Bacillus subtilis subsp.subtilis in solid state fermentation and its utility in improving food nutrition.Protein Pept. Lett. 28 : 1083-1089. - Yaver E. 2023. Dephytinized flaxseed flours by phytase enzyme and fermentation: functional ingredients to enhance the nutritional quality of noodles.
J. Sci. Food Agric. 103 : 1946-1953. - Pragya P, Sharma KK, Kumar S, Manisha F, Singh D, Kumar V, Singh B. 2023. Enhanced production and immobilization of phytase from
Aspergillus oryzae : a safe and ideal food supplement for improving nutrition.Lett. Appl. Microbiol. 76 : ovac077. - Thakur N, Patel SKS, Kumar P, Singh A, Devi N, Sandeep K,
et al . 2022. Bioprocess for hyperactive thermotolerantAspergillus fumigatus phytase and its application in dephytinization of wheat flour.Catal. Lett. 152 : 3220-3232. - Truelock CN, Yoder AD, Evans CE, Stark CR, Paulk CB. 2022. The effects of pelleting process parameters and phytase source on the in-feed stability of phytase.
Anim. Feed Sci. Technol. 294 : 115407. - Shivange AV, Schwaneberg U. 2017. Recent advances in directed phytase evolution and rational phytase engineering, pp. 145-172.
In Alcalde M (ed.),Directed Enzyme Evolution: Advances and Applications . Springer, NY, USA. - Herrmann KR, Hofmann I, Jungherz D, Wittwer M, Infanzón B, Hamer SN,
et al . 2021. Generation of phytase chimeras with low sequence identities and improved thermal stability.J. Biotechnol. 339 : 14-21. - Xing H, Wang P, Yan X, Yang Y, Li X, Liu R,
et al . 2023. Thermostability enhancement ofEscherichia coli phytase by error-prone polymerase chain reaction (epPCR) and site-directed mutagenesis.Front. Bioeng. Biotechnol. 11 : 1167530. - Yang LL, Shi HL, Liu F, Wang Z, Chen KL, Chen WS,
et al . 2022. Gene cloning of a highly active phytase fromLactobacillus plantarum and further improving its catalytic activity and thermostability through protein engineering.Enzyme Microb. Technol. 156 : 109997. - Wang Q, Liu X, Tian J, Wang Y, Zhang H, Wang Y,
et al . 2022. Enhancing the thermostability of phytase to boiling point by evolutionguided design.Appl. Environ. Microbiol. 88 : e00506-22. - Navone L, Vogl T, Luangthongkam P, Blinco JA, Luna-Flores CH, Chen X,
et al . 2021. Disulfide bond engineering of AppA phytase for increased thermostability requires co-expression of protein disulfide isomerase inPichia pastoris .Biotechnol. Biofuels 14 : 80. - Fakhravar A, Hesampour A. 2018. Rational design-based engineering of a thermostable phytase by site-directed mutagenesis.
Mol. Biol. Rep. 45 : 2053-2061. - Zhang Z, Yang J, Xie P, Gao Y, Bai J, Zhang C,
et al . 2022. Characterization of a thermostable phytase fromBacillus licheniformis WHU and further stabilization of the enzyme through disulfide bond engineering.Enzyme Microb. Technol. 142 : 109679. - Sanchez-Romero I, Ariza A, Wilson KS, Skjøt M, Vind J, De Maria L,
et al . 2013. Mechanism of protein kinetic stabilization by engineered disulfide crosslinks.PLoS One 8 : e70013. - Li J, Li X, Gai Y, Sun Y, Zhang D. 2019. Evolution of
E. coli phytase for increased thermostability guided by rational parameters.J. Microbiol. Biotechnol. 29 : 419-428. - Pang B, Zhou L, Cui W, Liu Z, Zhou Z. 2020. Improvement of the thermostability and activity of pullulanase from
Anoxybacillus sp. WB42.Appl. Biochem. Biotechnol. 191 : 942-954. - Huang H, Luo H, Yang P, Meng K, Wang Y, Yuan T,
et al . 2006. A novel phytase with preferable characteristics fromYersinia intermedia .Biochem. Biophys. Res. Commun. 350 : 884-889. - Yuan L, Huang Z, Zeng J, Guo JJ, Zhang T, Lü J. 2018. Fusion of phytase YiAPPA with the raw-starch binding domain and characterization of the fusion enzyme.
Biotechnol. Bullet. 34 : 200-207. - Waterhouse A, Bertoni M, Bienert S, Studer G, Tauriello G, Gumienny R,
et al . 2018. SWISS-MODEL: homology modelling of protein structures and complexes.Nucleic Acids Res. 46 : W296-W303. - Nezhad NG, Raja Abd Rahman RNZ, Normi YM, Oslan SN, Shariff FM, Leow TC. 2020. Integrative structural and computational biology of phytases for the animal feed industry.
Catalysts 10 : 844. - Fu D, Huang H, Luo H, Wang Y, Yang P, Mung K,
et al . 2008. A highly pH-stable phytase fromYersinia kristeensenii : cloning, expression, and characterization.Enzyme Microb. Technol. 42 : 499-505. - Anagnostopoulos C, Spizizen J. 1961. Requirements for transformation in
Bacillus subtilis .J. Bacteriol. 81 : 741. - Green MR, Sambrook J. 2012.
Molecular cloning: a laboratory manual , pp. 101-200. Cold Spring Harbor Laboratory Press, New York. - Laemmli UK. 1970. Cleavage of structural proteins during the assembly of the head of bacteriophage T4.
Nature 227 : 680-685. - Busto MD, Apenten RKO, Robinson DS, Wu Z, Casey R, Hughes RK. 1999. Kinetics of thermal inactivation of pea seed lipoxygenases and the effect of additives on their thermostability.
Food Chem. 65 : 323-329. - Lim JH, Hwang KY, Choi J, Lee DY, Ahn BY, Cho Y,
et al . 2001. Mutational effects on thermostable superoxide dismutase fromAquifex pyrophilus : understanding the molecular basis of protein thermostability.Biochem. Biophys. Res. Commun. 288 : 263-268. - Van Der Spoel D, Lindahl E, Hess B, Groenhof G, Mark AE, Berendsen HJ. 2005. GROMACS: fast, flexible, and free.
J. Comput. Chem. 26 : 1701-1718. - Morris GM, Huey R, Lindstrom W, Sanner MF, Belew RK, Goodsell DS,
et al . 2009. AutoDock4 and AutoDockTools4: automated docking with selective receptor flexibility.J. Comput. Chem. 30 : 2785-2791. - Wang CH, Nguyen PH, Pham K, Huynh D, Le TBN, Wang HL,
et al . 2016. Calculating protein-ligand binding affinities with MMPBSA: method and error analysis.J. Comput. Chem. 37 : 2436-2446. - Singh B, Satyanarayana T. 2006. Phytase production by thermophilic mold
Sporotrichum thermophile in solid-state fermentation and its application in dephytinization of sesame oil cake.Appl. Biochem. Biotechnol. 133 : 239-250. - Cai T, Cao J, Qiu S, Lyu C, Fan F, Hu S,
et al . 2023. Semi-rational evolution of ω-transaminase fromAspergillus terreus for enhancing the thermostability.Chinese J. Biotechnol. 39 : 2126-2140. - Zhang JW, Liu XQ, Tian J, LUO H, Yao B, Tu T. 2023. Improvement of the thermal stability of xylanase CbXyn10C from the thermophilic bacterium
Caldicellulosiruptor bescii based on structural information.Microbiol. China 50 : 5261-5274. - Dutta BS, Nordblad M, Woodley JM, Peters GH. 2016. A correlation between the activity of
Candida antarctica lipase B and differences in binding free energies of organic solvent and substrate.ACS Catal. 6 : 6350-6361.
Related articles in JMB
Article
Research article
J. Microbiol. Biotechnol. 2024; 34(8): 1660-1670
Published online August 28, 2024 https://doi.org/10.4014/jmb.2403.03031
Copyright © The Korean Society for Microbiology and Biotechnology.
Engineering of the Phytase YiAPPA to Improve Thermostability and Activity and Its Application Potential in Dephytinization of Food Ingredients
Jing Zeng*, Jianjun Guo, and Lin Yuan
Institute of Microbiology, Jiangxi Academy of Sciences, Nanchang 330096, Jiangxi Province, P.R. China
Correspondence to:Jing Zeng, zengjingwhu@126.com
Abstract
The aim of this study was to modify phytase YiAPPA via protein surficial residue mutation to obtain phytase mutants with improved thermostability and activity, enhancing its application potential in the food industry. First, homology modeling of YiAPPA was performed. By adopting the strategy of protein surficial residue mutation, the lysine (Lys) and glycine (Gly) residues on the protein surface were selected for site-directed mutagenesis to construct single-site mutants. Thermostability screening was performed to obtain mutants (K189R and K216R) with significantly elevated thermostability. The combined mutant K189R/K216R was constructed via beneficial mutation site stacking and characterized. Compared with those of YiAPPA, the half-life of K189R/K216R at 80°C was extended from 14.81 min to 23.35 min, half-inactivation temperature (T50 30) was increased from 55.12°C to 62.44°C, and Tm value was increased from 48.36°C to 53.18°C. Meanwhile, the specific activity of K189R/K216R at 37°C and pH 4.5 increased from 3960.81 to 4469.13 U/mg. Molecular structure modeling analysis and molecular dynamics simulation showed that new hydrogen bonds were introduced into K189R/K216R, improving the stability of certain structural units of the phytase and its thermostability. The enhanced activity was primarily attributed to reduced enzyme-substrate binding energy and shorter nucleophilic attack distance between the catalytic residue His28 and the phytate substrate. Additionally, the K189R/K216R mutant increased the hydrolysis efficiency of phytate in food ingredients by 1.73–2.36 times. This study established an effective method for the molecular modification of phytase thermostability and activity, providing the food industry with an efficient phytase for hydrolyzing phytate in food ingredients.
Keywords: Phytase, surficial residue, site-directed mutagenesis, thermostability, molecular dynamics simulation
Introduction
Phytate is the primary form of phosphorus in grasses and legumes [1, 2]. It can form stable and difficult-to-degrade complexes with proteins, minerals, and other substances in organisms, reducing the bioavailability of proteins and mineral elements [1, 2]. For this reason, phytate is considered an anti-nutritional factor. Phytase, also known as myo-inositol hexakisphosphate phosphohydrolase, can degrade phytate and generate inorganic phosphorus and phosphoinositides, eliminating the anti-nutritional effects of phytate [3, 4]. Owing to the lack of phytase in the human digestive system, phytate cannot be metabolized, which limits the absorption of nutrients in the intestine and reduces the nutritional value of food. Phytase can reduce the phytate level in food more effectively than cooking, boiling, fermentation, soaking, and other food processing methods. Adding phytase to plant-based food ingredients can degrade phytate and release inorganic phosphorus, which can improve the absorption and utilization of phosphorus in food, relieving the anti-nutritional effect of phytate [5-8]. Thus, phytase can be used as a new type of food additive in the food industry. However, numerous processing operations in the food industry necessitate execution under high temperatures. The limited thermostability of natural phytase poses a critical constraint in its practical utilization in food processing [9, 10]. Hence, the pursuit and advancement of improved thermostability of phytases remain a focal point of research among scholars worldwide.
Currently, protein engineering strategies used to enhance the thermostability of phytase include directed evolution (irrational design), semi-rational design, and rational design [11-14]. Rational design in the context of enzyme engineering involves directed modifications of enzymes based on the structure-activity relationship between molecular structure and function. Common methods employed in rational design include ion binding design, disulfide bond design, glycosylation site design, homology modeling, and molecular dynamics simulation [15-18]. For example, Sanchez-Romero
The strategy of protein surficial residue mutation was devised based on the above findings. Firstly, the GetArea software (http://curie.utmb.edu/getarea.html) was utilized to predict the localization of Gly and Lys residues on the protein surface. Next, each Gly residue was systematically substituted with Ala, and each Lys residue was replaced with Arg, leading to the construction of a mutant library. Directed mutagenesis was accomplished by carefully screening the resulting library to obtain mutants with specific modifications. Pang
The phytase YiAPPA (EC 3.1.3.2) derived from
Materials and Methods
Strains, Plasmid, and Culture Growth
Bioinformatic Analysis
The protein structure of the phytase YiAPPA was predicted using homology modeling by SWISS-MODEL (http://swissmodel.expasy.org) [24], with the molecular structure of the phytase YkAPPA from
Construction and Identification of Site-Specific Mutants
The primers were designed based on the base sequence of YiAPPA and the mutated amino acid sites by referring to the instructions of the QuickMutation kit, as shown in Table S1. Taking the construction of the site-directed mutant G16A as an example, the primers G16A-F and G16A-R were used to amplify linear fragments containing the vector sequence and the mutated gene sequence from the template recombinant plasmid pSTOP1622-
Induction, Expression, and Purification of Recombinant Phytase
The recombinant plasmid was transferred into competent
Determination of the Recombinant Phytase Activity
A 250-μl aliquot of the phytase solution (0.005 mg/ml) was diluted and thoroughly mixed with 750 μl of 0.25 mol/l sodium acetate buffer solution (pH 4.5). Thereafter, 2 ml of 1.5 mmol/l sodium phytate solution (0.25 mol/l sodium acetate buffer, pH 4.5) was added to the experimental group, and 2 ml of the end-point mixture (ammonium molybdate/ammonium vanadate/nitric acid) was added to the control group. The mixtures were thoroughly mixed and incubated at 37°C for 30 min. Subsequently, 2 ml of the end-point mixture was immediately added to the experimental group, and 2 ml of 1.5 mmol/l sodium phytate solution was added to the control group. All samples were thoroughly mixed and tested for their absorbance at 415 nm. The unit of phytase activity (U) was defined as the amount of phytase required to release 1 μmol/l of inorganic phosphorus from the 1.5 mmol/l sodium phytate solution per minute at pH 4.5 and 37°C.
Determination of the Enzymatic Properties of the Recombinant Phytase
The optimal reaction temperature, stability at 80°C, optimal reaction pH, stability at different pH levels, and protease resistance of the recombinant phytase were determined as previously reported [23].
The reaction system mentioned above was subjected to a 30-min reaction at temperatures ranging from 30°C to 90°C. The enzymatic activity of the samples was measured at various temperatures. The highest enzymatic activity was defined as 100%, and relative enzymatic activity was calculated for each temperature. A graph was plotted with relative enzymatic activity against temperature to determine the optimal reaction temperature. After incubating the enzyme solution at 80°C for 0–40 min, the enzymatic activity of the samples was determined according to the reaction system described above. The residual activity of the other samples was calculated using the enzymatic activity of the untreated sample as 100% and was plotted against temperature to evaluate the stability of the enzyme at 80°C.
The enzymatic activity at different pH levels (1.0–8.0) was determined. The relative enzymatic activity was calculated using the highest value as 100% enzymatic activity and was plotted against temperature to determine the optimal reaction pH. The enzyme solution was diluted using buffers of various pH levels (1.0–12.0), followed by a 2-h treatment at 37°C. Subsequently, the samples were diluted using the buffer with the optimal pH. The enzymatic activity of each sample was then measured based on the reaction system described above. The residual activity of other samples was calculated by using the enzymatic activity of the untreated sample as 100%, and the residual activity-temperature graph was plotted to evaluate the stability of enzymes under different pH conditions.
Solutions of 0.1 mg/ml of pepsin (0.25 mol/l Gly-HCl buffer, pH 2.0) and trypsin (0.25 mol/l Tris-HCl buffer, pH 7.0) were separately prepared. According to a proteinase-to-phytase mass ratio of 1:10, pepsin and trypsin were added to the phytase solution. The mixtures were then incubated at 37°C for 2 h, and a proteinase inhibitor was added to stop the reaction. Subsequently, the samples treated with proteinase were diluted 100-fold using the optimal pH buffer (0.25 mol/l sodium acetate buffer, pH 4.5). The enzymatic activity of the phytase before and after treatment with proteinase was determined using the reaction system described above.
Determination of Thermodynamic Parameters of the Recombinant Phytase
The thermodynamic parameter characterization of recombinant phytase mainly includes the determination of its half-life (
After incubating the enzyme solution (0.005 mg/ml) at 45°C, 50°C, 55°C, 60°C, 65°C, and 70°C for 30 min, the enzymatic activity of the samples was measured. Thereafter, the residual activity of the samples incubated at various temperatures was calculated with the enzymatic activity of the untreated sample as 100%. The temperature at which the residual activity was 50% was regarded as the
The
Determination of Kinetic Constant of the Recombinant Phytase
Sodium phytate solutions of various concentrations (0.0625%, 0.1%, 0.125%, 0.2%, 0.25%, 0.5%, 1.0%, and 1.5%) were prepared by diluting the stock solution with 0.25 mol/L sodium acetate buffer solution (pH 4.5). Equal volumes of phytase solution were added to each sodium phytate solution, and the enzymatic activity was determined following the method described above. Based on the double reciprocal plot, a graph was constructed with the reciprocal of substrate concentration as the x-axis and the reciprocal of enzyme activity as the y-axis. The slope of the line was
Molecular Dynamics Simulation and Molecular Docking
Molecular dynamics simulations were performed using Gromacs 2019.6 [32] at a temperature of 353 K (80°C) for a duration of 100 ns to analyze the thermodynamic fluctuations of the recombinant phytase. The simulations were conducted under constant temperature and pressure using the Amber14SB all-atom force field and the TIP3P water model. Periodic boundary conditions were applied during the simulations. The RMSD file reflecting the structure rigidity of the phytase molecule and the RMSF file reflecting the fluctuation of amino acid residues in phytase were obtained after the operation.
The recombinant phytase and phytate substrate were subjected to molecular docking using the software AutoDock 4.2.6 [33]. The binding energy and different energy contributions of the recombinant phytase and phytate substrate were calculated using the MMPBSA method [34]. Additionally, the software PyMOL v0.99 was employed to analyze the structural variations of the recombinant phytase at crucial amino acid residues.
Dephytinization of Food Ingredients
A 50-ml aliquot of 0.25 mol/l sodium acetate buffer (pH 4.5) was added to 5 g of food ingredients (wheat, corn, and soybean flours). After thoroughly mixing the samples, 10 U of purified recombinant phytase was added to the mixture. The mixture was then incubated at 55°C and 100 rpm for 4 h. Aliquots of samples were collected at 0.5, 1, and 2 h, and tested for the content of inorganic phosphorus and phytate as previously reported [35].
Statistical Analysis
Each of the enzymatic property tests was conducted in triplicate. Statistical analysis of experimental data was carried out using SigmaPlot 12.5, and all data are expressed as mean ± standard deviation.
Results and Discussion
Construction and Screening of Single-Site Mutants
This study intended to improve the thermostability of the phytase YiAPPA using protein surficial residue mutation. According to this molecular engineering strategy, the Lys and Gly residues located on the surface of YiAPPA were first analyzed. The amino acid sequence of YiAPPA was submitted to SWISS-MODEL for homology modeling using the molecular structure of the phytase YkAPPA (PDB ID: 4ARV) as the template, obtaining the molecular structure model for YiAPPA. The molecular structure model of YiAPPA was then submitted to the online prediction software GetArea. According to the working principle of GetArea, amino acid residues with a surface exposure ratio value greater than 50% were considered to be located on the surface of the protein molecule. The surficial Lys and Gly residues of YiAPPA predicted by GetArea are displayed in Table 1.
-
Table 1 . Surficial Lys and Gly residues of YiAPPA predicted using GetArea..
Amino acid Surface exposure ratio (%) Amino acid Surface exposure ratio (%) Gly16 100.0 Lys171 97.2 Lys35 92.9 Lys189 73.0 Gly81 72.7 Gly195 85.7 Gly86 67.3 Lys196 69.8 Gly117 100.0 Lys207 77.5 Gly119 77.3 Lys212 100.0 Lys121 79.7 Lys216 74.6 Lys129 96.4 Lys378 97.6 Lys130 68.1 Lys382 86.9 Lys144 70.8 Gly395 71.2 Lys152 72.2 Gly400 75.7 Gly159 81.2 Lys412 74.0 Lys167 99.4
Based on the prediction results in Table 1, YiAPPA mutants by replacing surface Lys with Arg (K35R, K121R, K129R, K130R, K144R, K152R, K167R, K171R, K189R, K196R, K207R, K212R, K216R, K378R, K382R, and K412R) and by replacing surface Gly with Ala (G16A, G81A, G86A, G117A, G119A, G159A, G195A, G395A, and G400A) were constructed. The heterologous recombination and expression of YiAPPA and its single-site mutants were carried out in
-
Figure 1. SDS-PAGE analysis and enzymatic activity and thermostability determination of the recombinant phytase YiAPPA and its single-site mutants.
(A) M. Protein marker; YiAPPA, G16A, K35R, G81A, G86A, G117A, G119A, K121R, K129R, K130R, K144R, K152R, G159A, K167R, K171R, K189R, G195A, K196R, K207R, K212R, K216R, K378R, K382R, G395A, G400A, K412R, and K189R/K216R are purified recombinant phytases.
Fig. 1B shows that the specific activity of recombinant YiAPPA was 3960.81 U/mg. The specific activity of the K216R mutant was 4352.96 U/mg, which was 1.10 times higher than that of recombinant YiAPPA. The specific activity of the K35R, K129R, K130R, and K144R mutants was significantly lower than that of YiAPPA. The specific activity of the other mutants was almost the same as that of recombinant YiAPPA. The residual activity of the mutants K189R and K216R after 20 min at 80°C was 43.00% and 46.00%, respectively, significantly higher than that of recombinant YiAPPA (33.99%). The above results demonstrated that the molecular engineering of the phytase YiAPPA through protein surficial residue mutation generated two mutants: K189R, with significantly improved thermostability, and K216R, with enhanced enzymatic activity and thermostability.
Construction and Enzymatic Property Analysis of K189R/K216R
Construction of K189R/K216R and the Preliminary Characterization of Its Enzymatic Properties
Mutation stacking is a method to improve enzymatic properties by combining beneficial mutations [36, 37]. Based on the previous study results, in this study, two mutants with significantly enhanced thermostability were stacked, constructing a combined mutant K189R/K216R. The combined mutant K189R/K216R was introduced into
-
Table 2 . Specific activity (37°C), kinetic parameters (37°C), thermodynamic parameters, and protease resistance of the recombinant phytase..
Recombinant phytase Specific activity (U/mg) Km (μM)kcat (s-1)kcat /Km (μM-1 s-1)55°C t 1/2 (min)80°C t 1/2 (min)T 50 30 (°C)Tm (°C)Pepsin resistance (%) Trypsin resistance (%) YiAPPA 3960.81 ± 105.34 126.24 ± 14.18 10375.27 ± 1256.73 82.19 30.09 ± 0.19 14.81 ± 0.37 55.12 ± 0.87 48.36 ± 0.53 84.17 ± 3.20 82.28 ± 4.18 K189R/K216R 4469.13 ± 154.19 111.59 ± 11.03 10363.29 ± 987.56 92.87 60.45 ± 0.63 23.35 ± 0.57 62.44 ± 0.76 53.18 ± 0.62 85.43 ± 4.09 83.97 ± 3.57 K189R 3979.56 ± 121.00 125.30 ± 13.09 10301.32 ± 1107.83 82.21 47.34 ± 0.38 17.86 ± 0.51 57.32 ± 0.64 49.93 ± 0.32 84.36 ± 2.98 83.01 ± 5.03 K216R 4352.96 ± 121.15 115.01 ± 9.86 10381.09 ± 1091.32 90.26 50.09 ± 0.52 18.90 ± 0.43 59.85 ± 0.73 50.79 ± 0.57 84.91 ± 3.25 83.30 ± 4.18
Determination of the Optimal Reaction Temperature and Thermodynamic Parameters for K189R/K216R
This study determined the optimal reaction temperature and thermodynamic parameters of the recombinant phytase YiAPPA and its mutant K189R/K216R, and the results are shown in Fig. 2 and Table 2. Fig. 2A shows that the optimal reaction temperature of the recombinant phytase YiAPPA and K189R/K216R mutant was 55°C. The relative activity of YiAPPA was similar to that of K189R/K216R within the range of 30°C–55°C. However, at 60°C–90°C, the relative activity of K189R/K216R was significantly higher than that of YiAPPA. The above results indicate that K189R/K216R retained higher relative activity within the higher temperature range, and this may be associated with its superior thermostability. Fig. 2B reveals that the residual activity of K189R/K216R was significantly higher than that of YiAPPA after incubation at 55°C or 80°C, indicating its significantly higher stability at 55°C and 80°C than YiAPPA.
-
Figure 2. Effect of temperature and pH on the activity and stability of the recombinant phytase.
(A) The effect of temperature on the activity of the recombinant phytase. (B) The effect of temperature on the stability of the recombinant phytase. (C) The effect of pH on the activity of the recombinant phytase. (D) The effect of pH on the stability of the recombinant phytase.
Based on these investigations, the thermodynamic parameter characterization of the recombinant phytase was carried out by determining the
Determination of the Optimal Reaction pH and pH Stability of K189R/K216R
This study determined the optimal reaction pH and the pH stability of the recombinant phytase YiAPPA and its mutant K189R/K216R, and the results are shown in Fig. 2C and 2D. The optimal reaction pH of the recombinant phytase YiAPPA and K189R/K216R mutant was 4.5, and they yielded similar relative enzyme activities at pH 1.0–8.0 (Fig. 2C). Within the pH range of 1.0–12.0, the stability of the recombinant phytase YiAPPA and K189R/K216R mutant was also similar (Fig. 2D). Within the pH range of 1.0–3.0, both exhibited over 80% residual enzyme activity. Within the pH range of 4.0–10.0, both demonstrated over 90% relative enzyme activity. However, within the pH range of 11.0–12.0, both were unstable and rapidly inactivated.
Determination of the Kinetic Parameters of K189R/K216R
The specific activity,
Determination of Protease Resistance of K189R/K216R
The results of the protease resistance analysis of the K189R/K216R mutant are shown in Table 2. The recombinant phytase YiAPPA and its mutants K189R, K216R, and K189R/K216R showed comparable protease resistance. Following pepsin treatment, recombinant YiAPPA retained 84.17% residual enzymatic activity, while the mutants K189R, K216R, and K189R/K216R retained 84.36%, 84.91%, and 85.43% residual enzymatic activity, respectively. After trypsin treatment, recombinant YiAPPA retained 82.28% residual enzymatic activity, whereas the mutants K189R, K216R, and K189R/K216R retained 83.01%, 83.30%, and 83.97% residual enzymatic activity, respectively. Overall, the introduction of mutations (K189R and/or K216R) in YiAPPA did not significantly affect its resistance to proteases.
Exploration of the Molecular Mechanism of Increased Thermostability and Activity of K189R/K216R
Molecular dynamics simulation of the recombinant phytase. At 353 K, a 100-ns molecular dynamics simulation was performed on the tertiary structure of YiAPPA and K189R/K216R to analyze their thermodynamic fluctuations, and the overall structural rigidity of the enzyme molecule and the fluctuations of amino acid residues were analyzed. The results are shown in Fig. 3. Fig. 3A shows that the RMSD value of YiAPPA and K189R/K216R fluctuated less after 40 ns and reached equilibrium, with the mean RMSD value of YiAPPA being 0.18 Å and that of K189R/K216R being 0.17 Å, that is, the mean RMSD value of the mutant K189R/K216R was slightly lower than that of YiAPPA. The RMSD value of an enzyme depends on its overall structural rigidity and stability. The lower the RMSD value, the greater the structural rigidity of the enzyme molecule, showing that it is more stable. The K189R/K216R mutant has a smaller RMSD value than YiAPPA, indicating it has a more stable molecular structure.
-
Figure 3. Molecular dynamics simulation results of the recombinant phytase.
(A) RMSD value analysis of the recombinant phytase YiAPPA and K189R/K216R at 353 K (80°C). (B) RMSF value analysis of the recombinant phytase YiAPPA and K189R/K216R at 353 K (80°C).
Fig. 3B presents the RMSF analysis results of YiAPPA and K189R/K216R. At Lys-189 and its nearby residues (serine (Ser)-190 to cysteine (Cys)-198, S190-C198), the RMSF value of K189R/K216R decreased compared with that of YiAPPA, showing that the surrounding amino acid residues became more stable due to the mutation of Lys-189 in YiAPPA. This finding means that this single-site mutation also improved the stability of amino acid residues at nearby sites. In the region adjacent to the mutation site of Lys-216 (K216) (threonine (Thr)-209 to Cys-214, T209-C214, valine (Val)-217 to leucine (Leu)-219, V217-L219), the RMSF value of K189R/K216R was significantly lower than that of YiAPPA. However, the RMSF of the K216 site was unchanged, indicating that although Lys mutation at position 216 in YiAPPA, as an essential mutation site, did not alter its RMSF, it improved the stability of amino acid residues in the nearby region.
These findings demonstrate that K189R/K216R had a slightly lower average RMSD value than YiAPPA, and the RMSF value of its mutation and nearby sites was significantly reduced. The molecular structure of K189R/K216R was more rigid, and some of its amino acid residues were more stable, leading to the higher thermostability of K189R/K216R than that of YiAPPA.
Molecular structure analysis of the recombinant phytase. Using the structure of the phytase YkAPPA (PDB ID 4ARV) as the template, the structure models of YiAPPA and K189R/K216R were constructed through homologous modeling, and the interaction forces between the amino acid residues of mutation sites and their surrounding sites before and after mutation were analyzed to explore the reasons for the increased rigidity of K189R/K216R. The results are shown in Fig. 4A and 4B. Lys-189 is located in the α-helix away from the catalytic pocket and has a long and flexible side chain. When Lys-189 was mutated to Arg, a hydrogen bond was formed between the newly added Lys at position 189 and the Ala at position 184. Based on the results of the RMSF analysis, it can be concluded that the mutation of Lys to Arg at position 189 in YiAPPA further stabilized both itself and the Ser at position 190 to Cys at position 198. This increase in hydrogen bonding between the newly added R189 (Arg-189) and A184 (Ala-184) enhanced the conformational stability of this protein region. As shown in Fig. 4A and 4B, Lys-216 is located in the loop structure near the catalytic pocket, and its mutation to Arg introduced a hydrogen bond between Arg-216 and Ser-218. The results of molecular dynamics simulation revealed significantly reduced RMSF values for Thr-209 to Cys-214 and Val-217 to Leu-219. Specifically, this decrease indicates enhanced rigidity in the loop structure (Thr-209 to Cys-214) and the β-sheet structure (Val-217 to Leu-219). Considering the RMSF result analysis, it is apparent that the newly formed hydrogen bond between R216 (Arg-216) and S218 (Ser-218) improved the stability of both the aforementioned loop structure and the β-sheet structure. Pang
-
Figure 4. Molecular structure analysis of the recombinant phytase.
(A) Molecular structure analysis of the recombinant phytase YiAPPA. (B) Molecular structure analysis of the recombinant phytase K189R/K216R. The blue parts in the figure represent the amino acid residues constituting the active site of the phytase. The red parts indicate the mutated amino acid residues (Lys-189 mutated to Arg-189; Lys-216 mutated to Arg-216) and the amino acid residues forming new hydrogen bonds with them (Ala-184 forms a new hydrogen bond with Arg-189; Ser-218 forms a new hydrogen bond with Arg-216). Hydrogen bonds are outlined in the magenta dashed box.
To further explore the molecular mechanism underlying the increased enzymatic activity of the mutant, the MMPBSA method was employed to calculate and decompose the total binding free energy of recombinant phytase and substrate phytate. Solvation energy refers to the sum of polar and non-polar solvation energy. The solvation energy values of YiAPPA and its mutants K189R, K216R, and K189R/K216R were 2055.79, 2072.31, 2098.97, and 2296.54 kcal/mol, respectively (Table 3). Theoretically, higher solvation energy is not conducive to substrate and enzyme binding [38]. The solvation energy of K216R and K189R/K216R, which displayed significantly improved enzymatic activity, was greater than that of YiAPPA. However, the absolute values of van der Waals energy and electrostatic energy for the two mutants were higher than those of YiAPPA, which was sufficient to offset the adverse effects of solvation during enzyme-substrate binding. The total binding energy of the K189R mutant was -67.29 kcal/mol, similar to that of YiAPPA. The total binding energy of the K216R mutant was -92.07 kcal/mol, which decreased by 43.41% compared to that of YiAPPA. The total binding energy of the K189R/K216R mutant was -97.07 kcal/mol, which was reduced by 51.20% relative to YiAPPA. These results show that the mutants K216R and K189R/K216R have increased stability and affinity after substrate binding, which is consistent with the changing trend of
-
Table 3 . Binding energy and various energy contributions of recombinant phytase and substrate phytate (kcal/mol)..
Recombinant phytase ΔGvdW ΔGele ΔGpolar ΔGnon-polar ΔGbind WT -18.35 ± 2.70 -2101.65 ± 13.67 2059.44 ± 9.03 -3.65 ± 0.04 -64.20 ± 16.60 K189R -21.96 ± 2.81 -2117.64 ± 7.96 2075.72 ± 2.09 -3.41 ± 0.08 -67.29 ± 8.69 K216R -18.59 ± 2.53 -2172.44 ± 8.28 2103.10 ± 2.77 -4.13 ± 0.03 -92.07 ± 9.09 K189R/K216R -19.51 ± 1.93 -2374.11 ± 8.52 2300.54 ± 15.18 -4.00 ± 0.02 -97.07 ± 17.51
-
Figure 5. The nucleophilic attack distance between His28 in recombinant phytase and phytate substrate.
(A) YiAPPA. (B) K189R. (C) K216R. (D) K189R/K216R. The red dotted line indicates atomic distance.
Application of the Recombinant Phytase as a Biocatalyst in Dephytinization of Food Ingredients
The phytate in food ingredients (wheat, corn, and soybean flours) was hydrolyzed by the recombinant phytase at 55°C. After hydrolysis, the content of inorganic phosphorus in the sample and the phytate degradation rates were measured and shown in Table 4. After the treatment of recombinant phytase YiAPPA and its K189R/K216R mutant, the content of inorganic phosphorus in the food ingredients and its phytate degradation rate were enhanced. In addition, compared with that of YiAPPA, the phytate hydrolysis efficiency of K189R/K216R (based on the inorganic phosphorus content in samples) against the food ingredients was increased by 2.36 (wheat flour), 2.28 (corn flour), and 1.73 times (soybean flour), respectively. After treatment with K189R/K216R at 55°C for 2 h, the inorganic phosphorus content in the wheat, corn, and soybean flour samples was 20.14 (wheat flour), 13.04 (corn flour), and 27.54 mg/g (soybean flour) with phytate degradation rates of 79.12%, 57.59%, and 81.52%, respectively. The above results reveal that the K189R/K216R mutant could more effectively reduce phytate in food ingredients and release inorganic phosphorus than YiAPPA, eliminating the anti-nutritional effect of phytate and improving the absorption and utilization of phosphorus in food.
-
Table 4 . Hydrolysis of phytates in food ingredients by the recombinant phytase..
Food Ingredient Recombinant Phytase Inorganic phosphorus (mg/g) Phytate degradation rate (%) 0.5 h 1 h 2 h 0.5 h 1 h 2 h Wheat flour YiAPPA 5.09 ± 0.39 7.03 ± 0.28 8.52 ± 0.67 19.99 27.62 33.47 K189R/K216R 6.65 ± 0.53 15.05 ± 0.56 20.14 ± 0.98 26.13 59.13 79.12 Corn flour YiAPPA 3.49 ± 0.36 4.89 ± 0.42 5.72 ± 0.37 15.41 21.60 25.26 K189R/K216R 4.58 ± 0.39 9.92 ± 0.51 13.04 ± 0.49 20.23 43.81 57.59 Soybean flour YiAPPA 8.94 ± 0.38 14.08 ± 0.51 15.95 ± 0.63 26.46 41.68 47.21 K189R/K216R 10.08 ± 0.59 21.31 ± 0.74 27.54 ± 0.66 29.84 63.08 81.52
This study successfully created the phytase mutant K189R/K216R with improved thermostability through protein surficial residue mutation. The
Supplemental Materials
Acknowledgments
This work was financially supported by grants from the National Natural Science Foundation of China (grant number 32160579) and Jiangxi Natural Science Foundation (grant numbers 20212BCJ23033, S2021GDQN2403, and 2022YJC2006).
Conflict of Interest
The authors have no financial conflicts of interest to declare.
Fig 1.
Fig 2.
Fig 3.
Fig 4.
Fig 5.
-
Table 1 . Surficial Lys and Gly residues of YiAPPA predicted using GetArea..
Amino acid Surface exposure ratio (%) Amino acid Surface exposure ratio (%) Gly16 100.0 Lys171 97.2 Lys35 92.9 Lys189 73.0 Gly81 72.7 Gly195 85.7 Gly86 67.3 Lys196 69.8 Gly117 100.0 Lys207 77.5 Gly119 77.3 Lys212 100.0 Lys121 79.7 Lys216 74.6 Lys129 96.4 Lys378 97.6 Lys130 68.1 Lys382 86.9 Lys144 70.8 Gly395 71.2 Lys152 72.2 Gly400 75.7 Gly159 81.2 Lys412 74.0 Lys167 99.4
-
Table 2 . Specific activity (37°C), kinetic parameters (37°C), thermodynamic parameters, and protease resistance of the recombinant phytase..
Recombinant phytase Specific activity (U/mg) Km (μM)kcat (s-1)kcat /Km (μM-1 s-1)55°C t 1/2 (min)80°C t 1/2 (min)T 50 30 (°C)Tm (°C)Pepsin resistance (%) Trypsin resistance (%) YiAPPA 3960.81 ± 105.34 126.24 ± 14.18 10375.27 ± 1256.73 82.19 30.09 ± 0.19 14.81 ± 0.37 55.12 ± 0.87 48.36 ± 0.53 84.17 ± 3.20 82.28 ± 4.18 K189R/K216R 4469.13 ± 154.19 111.59 ± 11.03 10363.29 ± 987.56 92.87 60.45 ± 0.63 23.35 ± 0.57 62.44 ± 0.76 53.18 ± 0.62 85.43 ± 4.09 83.97 ± 3.57 K189R 3979.56 ± 121.00 125.30 ± 13.09 10301.32 ± 1107.83 82.21 47.34 ± 0.38 17.86 ± 0.51 57.32 ± 0.64 49.93 ± 0.32 84.36 ± 2.98 83.01 ± 5.03 K216R 4352.96 ± 121.15 115.01 ± 9.86 10381.09 ± 1091.32 90.26 50.09 ± 0.52 18.90 ± 0.43 59.85 ± 0.73 50.79 ± 0.57 84.91 ± 3.25 83.30 ± 4.18
-
Table 3 . Binding energy and various energy contributions of recombinant phytase and substrate phytate (kcal/mol)..
Recombinant phytase ΔGvdW ΔGele ΔGpolar ΔGnon-polar ΔGbind WT -18.35 ± 2.70 -2101.65 ± 13.67 2059.44 ± 9.03 -3.65 ± 0.04 -64.20 ± 16.60 K189R -21.96 ± 2.81 -2117.64 ± 7.96 2075.72 ± 2.09 -3.41 ± 0.08 -67.29 ± 8.69 K216R -18.59 ± 2.53 -2172.44 ± 8.28 2103.10 ± 2.77 -4.13 ± 0.03 -92.07 ± 9.09 K189R/K216R -19.51 ± 1.93 -2374.11 ± 8.52 2300.54 ± 15.18 -4.00 ± 0.02 -97.07 ± 17.51
-
Table 4 . Hydrolysis of phytates in food ingredients by the recombinant phytase..
Food Ingredient Recombinant Phytase Inorganic phosphorus (mg/g) Phytate degradation rate (%) 0.5 h 1 h 2 h 0.5 h 1 h 2 h Wheat flour YiAPPA 5.09 ± 0.39 7.03 ± 0.28 8.52 ± 0.67 19.99 27.62 33.47 K189R/K216R 6.65 ± 0.53 15.05 ± 0.56 20.14 ± 0.98 26.13 59.13 79.12 Corn flour YiAPPA 3.49 ± 0.36 4.89 ± 0.42 5.72 ± 0.37 15.41 21.60 25.26 K189R/K216R 4.58 ± 0.39 9.92 ± 0.51 13.04 ± 0.49 20.23 43.81 57.59 Soybean flour YiAPPA 8.94 ± 0.38 14.08 ± 0.51 15.95 ± 0.63 26.46 41.68 47.21 K189R/K216R 10.08 ± 0.59 21.31 ± 0.74 27.54 ± 0.66 29.84 63.08 81.52
References
- Philippi H, Sommerfeld V, Windisch W, Olukosi OA, Monteiro A, Rodehutscord M. 2023. Interactions of zinc with phytate and phytase in the digestive tract of poultry and pigs: a review.
J. Sci. Food Agric. 103 : 7333-7342. - Handa V, Sharma D, Kaur A, Kumar Arya S. 2020. Biotechnological applications of microbial phytase and phytate in food and feed industries.
Biocatal. Agric. Biotechnol. 25 : 101600. - Hussain SM, Hanif S, Sharif A, Bashir F, Iqbal HMN. 2022. Unrevealing the sources and catalytic functions of phytase with multipurpose characteristics.
Catal. Lett. 152 : 1358-1371. - Rizwanuddin S, Kumar V, Naik B, Singh P, Mishra S, Rustagi S,
et al . 2023. Microbial phytase: their sources, production, and role in the enhancement of nutritional aspects of food and feed additives.J. Agric. Food Res. 12 : 100559. - Song HY, El Sheikha AF, Hu DM. 2019. The positive impacts of microbial phytase on its nutritional applications.
Trends Food Sci. Technol. 86 : 553-562. - Singh B, Kumar G, Kumar V, Singh D. 2021. Enhanced phytase production by
Bacillus subtilis subsp.subtilis in solid state fermentation and its utility in improving food nutrition.Protein Pept. Lett. 28 : 1083-1089. - Yaver E. 2023. Dephytinized flaxseed flours by phytase enzyme and fermentation: functional ingredients to enhance the nutritional quality of noodles.
J. Sci. Food Agric. 103 : 1946-1953. - Pragya P, Sharma KK, Kumar S, Manisha F, Singh D, Kumar V, Singh B. 2023. Enhanced production and immobilization of phytase from
Aspergillus oryzae : a safe and ideal food supplement for improving nutrition.Lett. Appl. Microbiol. 76 : ovac077. - Thakur N, Patel SKS, Kumar P, Singh A, Devi N, Sandeep K,
et al . 2022. Bioprocess for hyperactive thermotolerantAspergillus fumigatus phytase and its application in dephytinization of wheat flour.Catal. Lett. 152 : 3220-3232. - Truelock CN, Yoder AD, Evans CE, Stark CR, Paulk CB. 2022. The effects of pelleting process parameters and phytase source on the in-feed stability of phytase.
Anim. Feed Sci. Technol. 294 : 115407. - Shivange AV, Schwaneberg U. 2017. Recent advances in directed phytase evolution and rational phytase engineering, pp. 145-172.
In Alcalde M (ed.),Directed Enzyme Evolution: Advances and Applications . Springer, NY, USA. - Herrmann KR, Hofmann I, Jungherz D, Wittwer M, Infanzón B, Hamer SN,
et al . 2021. Generation of phytase chimeras with low sequence identities and improved thermal stability.J. Biotechnol. 339 : 14-21. - Xing H, Wang P, Yan X, Yang Y, Li X, Liu R,
et al . 2023. Thermostability enhancement ofEscherichia coli phytase by error-prone polymerase chain reaction (epPCR) and site-directed mutagenesis.Front. Bioeng. Biotechnol. 11 : 1167530. - Yang LL, Shi HL, Liu F, Wang Z, Chen KL, Chen WS,
et al . 2022. Gene cloning of a highly active phytase fromLactobacillus plantarum and further improving its catalytic activity and thermostability through protein engineering.Enzyme Microb. Technol. 156 : 109997. - Wang Q, Liu X, Tian J, Wang Y, Zhang H, Wang Y,
et al . 2022. Enhancing the thermostability of phytase to boiling point by evolutionguided design.Appl. Environ. Microbiol. 88 : e00506-22. - Navone L, Vogl T, Luangthongkam P, Blinco JA, Luna-Flores CH, Chen X,
et al . 2021. Disulfide bond engineering of AppA phytase for increased thermostability requires co-expression of protein disulfide isomerase inPichia pastoris .Biotechnol. Biofuels 14 : 80. - Fakhravar A, Hesampour A. 2018. Rational design-based engineering of a thermostable phytase by site-directed mutagenesis.
Mol. Biol. Rep. 45 : 2053-2061. - Zhang Z, Yang J, Xie P, Gao Y, Bai J, Zhang C,
et al . 2022. Characterization of a thermostable phytase fromBacillus licheniformis WHU and further stabilization of the enzyme through disulfide bond engineering.Enzyme Microb. Technol. 142 : 109679. - Sanchez-Romero I, Ariza A, Wilson KS, Skjøt M, Vind J, De Maria L,
et al . 2013. Mechanism of protein kinetic stabilization by engineered disulfide crosslinks.PLoS One 8 : e70013. - Li J, Li X, Gai Y, Sun Y, Zhang D. 2019. Evolution of
E. coli phytase for increased thermostability guided by rational parameters.J. Microbiol. Biotechnol. 29 : 419-428. - Pang B, Zhou L, Cui W, Liu Z, Zhou Z. 2020. Improvement of the thermostability and activity of pullulanase from
Anoxybacillus sp. WB42.Appl. Biochem. Biotechnol. 191 : 942-954. - Huang H, Luo H, Yang P, Meng K, Wang Y, Yuan T,
et al . 2006. A novel phytase with preferable characteristics fromYersinia intermedia .Biochem. Biophys. Res. Commun. 350 : 884-889. - Yuan L, Huang Z, Zeng J, Guo JJ, Zhang T, Lü J. 2018. Fusion of phytase YiAPPA with the raw-starch binding domain and characterization of the fusion enzyme.
Biotechnol. Bullet. 34 : 200-207. - Waterhouse A, Bertoni M, Bienert S, Studer G, Tauriello G, Gumienny R,
et al . 2018. SWISS-MODEL: homology modelling of protein structures and complexes.Nucleic Acids Res. 46 : W296-W303. - Nezhad NG, Raja Abd Rahman RNZ, Normi YM, Oslan SN, Shariff FM, Leow TC. 2020. Integrative structural and computational biology of phytases for the animal feed industry.
Catalysts 10 : 844. - Fu D, Huang H, Luo H, Wang Y, Yang P, Mung K,
et al . 2008. A highly pH-stable phytase fromYersinia kristeensenii : cloning, expression, and characterization.Enzyme Microb. Technol. 42 : 499-505. - Anagnostopoulos C, Spizizen J. 1961. Requirements for transformation in
Bacillus subtilis .J. Bacteriol. 81 : 741. - Green MR, Sambrook J. 2012.
Molecular cloning: a laboratory manual , pp. 101-200. Cold Spring Harbor Laboratory Press, New York. - Laemmli UK. 1970. Cleavage of structural proteins during the assembly of the head of bacteriophage T4.
Nature 227 : 680-685. - Busto MD, Apenten RKO, Robinson DS, Wu Z, Casey R, Hughes RK. 1999. Kinetics of thermal inactivation of pea seed lipoxygenases and the effect of additives on their thermostability.
Food Chem. 65 : 323-329. - Lim JH, Hwang KY, Choi J, Lee DY, Ahn BY, Cho Y,
et al . 2001. Mutational effects on thermostable superoxide dismutase fromAquifex pyrophilus : understanding the molecular basis of protein thermostability.Biochem. Biophys. Res. Commun. 288 : 263-268. - Van Der Spoel D, Lindahl E, Hess B, Groenhof G, Mark AE, Berendsen HJ. 2005. GROMACS: fast, flexible, and free.
J. Comput. Chem. 26 : 1701-1718. - Morris GM, Huey R, Lindstrom W, Sanner MF, Belew RK, Goodsell DS,
et al . 2009. AutoDock4 and AutoDockTools4: automated docking with selective receptor flexibility.J. Comput. Chem. 30 : 2785-2791. - Wang CH, Nguyen PH, Pham K, Huynh D, Le TBN, Wang HL,
et al . 2016. Calculating protein-ligand binding affinities with MMPBSA: method and error analysis.J. Comput. Chem. 37 : 2436-2446. - Singh B, Satyanarayana T. 2006. Phytase production by thermophilic mold
Sporotrichum thermophile in solid-state fermentation and its application in dephytinization of sesame oil cake.Appl. Biochem. Biotechnol. 133 : 239-250. - Cai T, Cao J, Qiu S, Lyu C, Fan F, Hu S,
et al . 2023. Semi-rational evolution of ω-transaminase fromAspergillus terreus for enhancing the thermostability.Chinese J. Biotechnol. 39 : 2126-2140. - Zhang JW, Liu XQ, Tian J, LUO H, Yao B, Tu T. 2023. Improvement of the thermal stability of xylanase CbXyn10C from the thermophilic bacterium
Caldicellulosiruptor bescii based on structural information.Microbiol. China 50 : 5261-5274. - Dutta BS, Nordblad M, Woodley JM, Peters GH. 2016. A correlation between the activity of
Candida antarctica lipase B and differences in binding free energies of organic solvent and substrate.ACS Catal. 6 : 6350-6361.