-
The enzyme flexible regions with high conformational flexibility are always used as a potential target for molecular modification in protein engineering. The flexible regions of CPA were mainly targeted by two methods (B-factor and RMSF). The B-factor is used to characterize the describe the attenuation of X-ray or neutron scattering caused by thermal motion. Where the higher the B-Factor value, the more flexible the enzyme conformation is[24]. As seen in Fig. 1, the average B-Factor value of residue was calculated by HotSpot Wizard 3.0, and a total of 16 highly flexible regions of CPA were identified.
Figure 1.
Average B-factor values of carboxypeptidase A. Flexible residues were located above the dotted line.
RMSF represents the amplitude of fluctuation of an atom relative to its average position in the simulated system, reflecting the local flexibility of protein conformation at the atomic level[21]. The higher the RMSF value, the more flexible the enzyme conformation is. As shown in Fig. 2, the RMSF values of CPA were calculated by molecular dynamics simulation, and a total of twelve flexible regions of CPA were screened.
Figure 2.
The RMSF values of carboxypeptidase A. Flexible residues were located above the dotted line.
To improve the prediction accuracy, B-factor and RMSF values were comprehensively analyzed. Finally, the regions of Ala1-Arg2, Asn5-Thr6, Gly55-Asn58, Gln92-Ser95, Val132-Leu137, Lys153-Ala154, Lys168-Tyr169, Lys231-Tyr234, Try248-Gln249, and Asn306-Asn307 were considered highly flexible regions, laying the foundation for the subsequent rigidification of flexible regions.
Selection of potential disulfide bonds in flexible regions of CPA
-
The introduction of disulfide bonds is always one of the strategies to stabilize the flexible regions. Potential residue pairs for introducing disulfide bonds in CPA were analyzed by two online programs, DbD and MODIP (Supplemental Fig. S1). There were 37 pairs of potential disulfide bonds in CPA predicted by DbD software. Besides, there were eight pairs of potential disulfide bonds predicted by the MODIP program, and they were categorized into Grade A, Grade B, Grade C and Grade D according to the likelihood of disulfide bond formation. To improve the prediction reliability of potential disulfide bonds in CPA, twenty potential disulfide bonds belonging to both DbD and MODIP classes A and B were screened. The residues including Asp93, Ser136 and Lys153 were located in these flexible regions. Therefore, three pairs of residues for introducing disulfide bonds Asp93-Phe96, Ser136-Pro160 and Lys153-Ser251 were initially screened for subsequent analysis.
Analysis of conserved residues of CPA
-
Enzymes not only need enough rigidity to maintain their stability, but also require adequate conformational flexibility to exert its catalytic activity. Thus, there is a trade-off between activity and stability, implying that an increase in stability is accompanied by a concomitant decrease in activity[25]. It was reported that residues associated with the catalytic function of enzymes were typically highly conserved[26]. A total of 151 conserved residues in the CPA sequence were determined by conservativeness analysis of PSSM, containing the screened residue Ser136 (Supplemental Fig. S2). Thus, the disulfide bond Ser136-Pro160 was excluded to prevent the catalytic activity of CPA from being adversely affected when engineering thermostability. Finally, the two CPA disulfide bonds Asp93-Phe96 and Lys153-Ser251 were retained and sequentially named as mutants D93C/F96C and K153C/S251C.
Homology Modeling of CPA mutants
-
Using the crystal structures of CPA (PDB ID:1M4L) as a template, three-dimensional structures of mutants D93C/F96C and K153C/S251C were obtained through SWISS-MODEL online server. The quality of the models was evaluated through the SAVES online platform. The results of Ramachandran plots showed that mutants D93C/F96C and K153C/S251C had more than 95% residues in the favorable regions (Supplemental Fig. S3). And, the results of ERRAT indicated that the overall quality factors of D93C/F96C and K153C/S251C were 95.2703 and 95.5932, respectively (Supplemental Fig. S4). Thus, the three-dimensional models of these two mutants were relatively reliable and closer to their real structures.
Evaluation of CPA mutants with potential enhanced thermostability
-
MD simulations were performed to explore the conformational stability of CPA structure. RMSD reflects the average amount of movement of backbone atoms throughout the whole protein structure, which is negatively correlated with the thermostability of the enzyme[21]. The average RMSD values of mutants D93C/F96C and K153C/S251C were 0.1386 nm (Fig. 3a) and 0.1136 nm (Fig. 3b), which were lower than that of WT (0.1416 nm). It suggested that mutations could make the overall conformation more rigid, contributing to enhancing the thermostability of CPA[27].
Figure 3.
The comparison of RMSD values between wild-type and mutant carboxypeptidase A. (a) WT and D93C/F96C. (b) WT and K153C/S251C.
The effect of mutation on the local conformational stability of CPA was explored by RMSF. Compared with WT, the mutants D93C/F96C (Fig. 4a) and K153C/S251C (Fig. 4b) both showed a decrease in RMSF values in mutant regions, showing that the mutations stabilize the local conformation of CPA. Taken together, the two mutants D93C/F96C and K153C/S251C possessed the potential to enhance the thermostability compared with WT. And, they were used for the subsequent experimental analysis of CPA thermostability screening.
Figure 4.
The comparison of RMSF values between wild-type and mutant carboxypeptidase A. (a) WT and D93C/F96C. (b) WT and K153C/S251C. The mutant regions are circled in dotted wireframes.
Construction, expression, and purification of recombinant CPA and its mutants
-
The plasmid pPIC9K contains the histidine dehydrogenase sequence (HIS4). It was shown that a large number of single colonies capable of synthesizing histidine grew on the MD plate (Supplemental Fig. S5). The result showed that the recombinant plasmid pPIC9K had been successfully transferred into host cells.
Because the plasmid pPIC9K contains the kanamycin resistant gene that allows P. pastoris to tolerate G418 disulfate salt. The result showed that the number of viable transformants gradually decreased as the concentration of genistein increased (Supplemental Fig. S6). To some extent, the higher the level of resistance to the genotoxin G418, the higher the copy number of the target gene, which may increase the level of enzyme expression. Therefore, the transformants growing on YPD plates with 4 mg·mL−1 G418 were selected for expression.
From the SDS-PAGE result of fermentation supernatants, there was a band appeared near 35 kDa, which was more similar to the theoretical molecular weight of CPA (Fig. 5a). The result preliminarily suggested that P. pastoris GS115 realized the heterologous expression of CPA and its mutants. To improve the purity of recombinant enzymes, hydrophobic interaction chromatography was used for purification. And, the SDS-PAGE of purified WT, D93C/F96C and K153C/S251C was shown in Fig. 5b. The purity of purified enzymes exceeded 90% detected by Image J. There were six histidine tags at the N-terminal of recombinant enzymes, the specific bands in the western-blot experiment once again proved that the purified enzymes with a molecular weight of 35 kDa was recombinant CPA and its mutants (Fig. 5c). The concentration of WT, D93C/F96C and K153C/S251 measured by Bradford's method were 0.321, 0.303, and 0.289 mg·mL−1, respectively.
Figure 5.
The SDS-PAGE analysis of (a) fermentation supernatants and (b) purified components expressed by P. pastoris. (c) The western blot analysis of purified components expressed by P. pastoris. M, protein marker (10−180 kDa). 1, WT. 2, D93C/F96C. 3, K153C/S251C. Each sample was prepared by boiling for 5 min and loaded at 20 μL per lane.
Enzymatic analysis of recombinant CPA and its mutants
-
To determine the OTA degradation efficiency of recombinant CPA and its mutants, the residual OTA in the reaction system was detected by HPLC. It can be seen in Fig. 6 that the peak time of the OTA standard was 4.150 min, and the peak time of the OTα standard was 3.205 min. Compared with the control, the peak area of samples treated with recombinant enzymes decreased significantly at 4.1 min, indicating that OTA was successfully degraded. A new chromatographic peak appeared around 3.2 min, and its retention time was consistent with that of the OTα standard. The result indicated that the OTA degradation product of WT, D93C/F96C and K153C/S251C was OTα.
It was found that CPA hydrolyzed the amide bond of OTA to generate OTα and L-phenylalanine[21], which was consistent with the principle of hydrolyzing Z-Phe-Leu. Given the toxicity of OTA, Z-Phe-leu was used as a substrate to study the enzymatic properties of CPA. The specific enzyme activity of recombinant WT, D93C/F96C and K153C/S251C was determined by the standard curve of L-leucine solution y = 0.0055x + 0.0857 (R2 = 0.995). The specific enzyme activity was 11.113 ± 0.298, 13.816 ± 0.511, and 10.107 ± 0.255 U·mg−1 for WT, D93C/F96C and K153C/S251C, respectively. The specific enzyme activity of D93C/F96C was increased by 24.32% compared with that of WT, indicating that the introduction of the disulfide bond at sites 93 and 96 did not adversely affect the activity. While the specific enzyme activity of the mutant K153C/S251C was reduced by 9.05% compared with that of WT, showing that the mutations at sites 153 and 251 adversely affected the enzymatic activity of CPA.
Thermostability measurements of recombinant CPA and its mutants
-
The reaction system of recombinant enzymes and substrate Z-Phe-Leu was processed at different temperatures for a certain period. And, the residual enzyme activity of recombinant enzymes was measured. As shown in Fig. 7a, the activity increased and then decreased with the increase of reaction temperature, and reached the highest at the optimal temperature. The maximum activity of WT at 40 °C indicated that the optimum temperature of WT was 40 °C. While the optimal temperatures of mutants D93C/F96C and K153/S251 were 10 °C higher than that of WT. The results showed that mutations improved the optimal temperature, which was conducive to the better catalytic activity of CPA at higher temperatures.
Figure 7.
Thermal stability of wild-type and mutant carboxypeptidase A. (a) The optimum temperature of wild-type and mutant carboxypeptidase A. (b) The half-life of wild-type and mutant carboxypeptidase A. (c) Half inactivation temperature of wild-type and mutant carboxypeptidase A.
The half-life (t1/2) represents the time required for the enzyme to lose half of its activity. The WT, D93C/F96C and K153C/S251C was subjected to treatment at 65 °C for different times, respectively. As shown in Fig. 7b, the half-life curve of WT, D93C/F96C and K153C/S251C was y = −0.05913x + 3.4361 (R2 = 0.98), y = −0.04598x + 3.6209 (R2 = 0.99) and y = 0.04898x + 3.3817 (R2 = 0.97), respectively. Since the slope of the half-life curve is the inactivation rate constant kd, the t1/2 value of WT at 65 °C can be calculated as 11.7 min. The t1/2 of D93C/F96C and K153C/S251C was 15.1 and 14.2 min, respectively, which were 3.4 and 2.5 min higher than that of WT. The above results showed that the heat resistance of CPA was enhanced after the introduction of the disulfide bond, and the activity could be maintained for a longer period at a higher temperature.
The half-inactivation temperature (T50) refers to the temperature at which the activity drops to 50% of the initial activity. In this study, the recombinant WT, D93C/F96C and K153C/S251C was first subjected to different temperatures for a certain period, respectively. From the half-inactivation temperature curves (Fig. 7c), it can be seen that the activity of WT and its mutants decreased with increasing temperature, but the activity of WT decreased more significantly. After 15 min of treatment at 70 °C, the activity of WT was only 35% of the initial activity, whereas the mutants D93C/F96C and K153C/S251C were able to maintain about 50% of the initial activity. The T5015 of WT was calculated to be 58.0 °C, while the T5015 of mutants D93C/F96C and K153C/S251C were 66.5 and 69.4 °C, which were 8.5 and 11.4 °C higher than that of WT, respectively. The above results indicated that mutants D93C/F96C and K153C/S251C were more stable and could maintain better activity at higher temperatures.
Kinetic parameter measurements of recombinant CPA and its mutants
-
The enzyme kinetic constants, Km and Kcat/Km, were measured to evaluate the enzymatic properties of WT and its mutants. The Michaelis constant Km reflects the affinity of the enzyme for the substrate. The smaller the value is, the better the affinity of the enzyme for the substrate. As shown in Table 1, the Km of D93C/F96C and K153C/S251 were not significantly different from those of WT, indicating that the mutation did not affect the affinity for the substrate. The Kcat/Km denotes the catalytic efficiency of the enzyme. And, the larger the value is, the higher the catalytic efficiency of the enzyme is. As shown in Table 1, the Kcat/Km of D93C/F96C was the largest, whose catalytic efficiency was increased by 43.15% compared with that of WT. While, the Kcat/Km value of K153C/S251 was decreased by 8.08%, demonstrating that its catalytic efficiency was slightly lower than that of WT.
Table 1. Kinetic parameters of wild-type and mutant carboxypeptidase A.
Enzyme Km (μM) Vmax (μM·min−1) Kcat/Km (μM−1·s−1) WT 0.277 ± 0.012 1.833 ± 0.014 4.549 × 10−3 D93C/F96C 0.271 ± 0.021 2.569 ± 0.036 6.512 × 10−3 K153C/S251C 0.268 ± 0.043 1.641 ± 0.047 4.209 × 10−3 The above results of recombinant WT, D93C/F96C and K153C/S251C were similar to those of specific enzyme activities. It indicated that D93C/F96C achieved a dual improvement in thermostability and activity. Zhou et al. also found that the stability and activity of LPMOs could be improved simultaneously[11]. While the catalytic activity of K153C/S251C suffered from undesired effects when engineering thermostability, exhibiting a prevalent phenomenon 'stability-activity trade-off'[25]. Ming et al. also reported a similar result that the OTA degradation ability of CPA mutant R124K and S134P decreased to varying degrees while improving their thermostability[21].
Advanced structure evaluation of recombinant CPA and its mutants
-
The circular dichroic absorption spectra of recombinant CPA at 200~260 nm were scanned to investigate the effect of mutations on CPA secondary structure. Subsequently, the percentage content of secondary structures such as α-helixes, β-strands, β-turns and coils were determined by CDpro computer software. As shown in Fig. 8, the contents of α-helixes, β-strands, β-turns, and coils for WT were 24.6%, 26.6%, 21.6%, and 27.5%, respectively. Compared with WT, the mutants D93C/F96C and K153C/S251C showed an increase in α-helixes by 5.1% and 10.7%, while a decrease in β-strands, β-turns and coils. The above results indicated that the increase of α-helixes might be a key cause of their structural preservation[19], which was conducive to the improvement of the thermostability of CPA.
Figure 8.
The contents of secondary structures (α-helix, β-strand, β-turn, and coil) in CPA and its mutants.
The fluorescence emission spectra of recombinant CPA at 290~500 nm was scanned to investigate the effect of mutation on CPA tertiary structure. As shown in Fig. 9, WT and its mutants D93C/F96C and K153C/S251C all had a maximum emission at around 335 nm, demonstrating the characteristics of tryptophan fluorescence[28]. It suggested that there was no obvious impact on the tertiary structure of CPA when introducing disulfide bonds. A similar result was reported in 1,4-α-glucan branching enzyme. Introducing disulfide bonds might only effect the secondary structure of it while the tertiary structure continued a similar trend[29].
Molecular interactions analysis of recombinant CPA and its mutants
-
The formation of disulfide bonds was quantified by detecting the amount of free sulfhydryl groups by DNTB[30]. As shown in Table 2, it was determined that WT possessed no natural disulfide bond. While, the mutants D93C/F96C and K153C/S251C both had one disulfide bond. The above results indicated that the mutants D93C/F96C and K153C/S251C successfully formed intramolecular disulfide bonds, realizing the purpose of enhancing the thermostability of CPA by introducing disulfide bonds.
Table 2. Comparison of disulfide bond number between wild-type and mutant carboxypeptidase A.
Enzyme A1 A2 Disulfide bond WT 0.456 ± 0.009 0.342 ± 0.010 0 D93C/F96C 1.434 ± 0.035 0.367 ± 0.018 1 K153C/S251C 1.549 ± 0.050 0.464 ± 0.023 1 The mutation of residues often leads to complex changes in multiple intramolecular interactions. In addition to analyzing disulfide bonds, ProteinTools online software was conducted to investigate the changes in the number of hydrogen bonds in mutant regions. Compared with WT, the mutant D93C/F96C showed an increase from one to two hydrogen bonds at mutation sites 93 and 96, increasing the thermostability by maintaining the rigidity of the enzyme structure (Fig. 10a). While, the mutant K153C/S251C lost the hydrogen bond formed between the mutation site 153 and 251, which might result in the decrease of enzyme activity (Fig. 10b). This was because hydrogen bonds were important factors for stabilizing the enzyme secondary structure[6].
Figure 10.
Comparison of the number of hydrogen bonds in mutant regions between wild-type and its mutant carboxypeptidase A. (a) WT and D93C/F96C. (b) WT and K153C/S251C.
The changes in surface charges for WT and its mutants were analyzed by PyMOL software. The enhanced thermostability of D93C/F96C and K153C/S251C might also be related to the surface charge redistribution in mutant regions. As shown in Fig. 11a, the WT was electrically neutral at the sites of Asp93 and Phe96, with the region of Asp93-Phe96 showing electronegative. Whereas residues at sites 93 and 96 were both mutated to Cys, the mutant D93C/F96C exhibited positive electronegativity at sites 93 and 96. And, the mutant region shifted from negative to positive electronegativity. The phenomenon might be caused by the reduction of a carboxyl group and the formation of two sulfhydryl groups in residue structures. For mutant K153C/S251C, when replacing positively charged Lys at site 153 and positively charged Ser at site 251 with uncharged Cys, not only did the surface charges of mutation sites change from positive to neutral, but also the surface charges of the mutation region Lys153-Ser251 showed a similar trend (Fig. 11b). The disappearance of an amino group and a hydroxyl group and the formation of two sulfhydryl groups could be associated with the changes in surface charge of K153C/S251C. Chen et al. also reported a similar result that the redistribution of surface electrostatic charges enhanced the thermostability of glycosyltransferase UGT76G1[30]. Besides, Arabnejad et al. found that the positive effect on thermostability of halohydrin dehalogenase D162T could be contributed to the redistribution of surface electrostatic charges caused by the removal of the carboxyl group[31].
-
All data generated or analyzed during this study are included in this published article.
-
About this article
Cite this article
Zhang H, Zhao Z, Zhu M, Logrieco AF, Wang H, et al. 2024. Enhancing the thermostability of carboxypeptidase A by rational design of disulfide bonds. Food Innovation and Advances 3(2): 191−201 doi: 10.48130/fia-0024-0017
Enhancing the thermostability of carboxypeptidase A by rational design of disulfide bonds
- Received: 14 March 2024
- Revised: 04 June 2024
- Accepted: 05 June 2024
- Published online: 25 June 2024
Abstract: Carboxypeptidase A(CPA) has a great potential application in the food and pharmaceutical industry due to its capability to hydrolyze ochratoxin A(OTA) and remove the bitterness of peptide. However, CPA is a mesophilic enzyme that cannot adequately exert its catalytic activity at elevated temperatures, which seriously restricts its industrial application. In this study, the rational design of disulfide bonds was introduced to improve the thermostability of CPA. The highly flexible regions of CPA were predicted through the HotSpot Wizard program and molecular dynamics (MD) simulations. Then, DbD and MODIP online servers were conducted to predict potential residue pairs for introducing disulfide bonds in CPA. After the conservativeness analysis of the PSSM matrix and the structural analysis of the MD simulation, two mutants with potentially enhanced thermostability were screened. Results showed that these mutants D93C/F96C and K153C/S251C compared to the wild-type(WT) exhibited increase by 10 and 10 °C in Topt, 3.4 and 2.7 min in t1/2 at 65 °C, in addition to rise of 8.5 and 11.4 °C in T5015, respectively. Furthermore, the molecular mechanism responsible for thermostability was investigated from the perspective of advanced structure and molecular interactions. The enhanced thermostability of both mutants was not only associated with the more stable secondary structure and the introduction of disulfide bonds but also related to the changes in hydrogen bonds and the redistribution of surface charges in mutant regions. This study showed for the first time that the rational design of disulfide bonds is an effective strategy to enhance the thermostability of CPA, providing in this way a broader industrial application.
-
Key words:
- Disulfide bonds /
- Thermostability /
- Carboxypeptidase A /
- Rational design