Next Article in Journal
Miglitol, an Oral Antidiabetic Drug, Downregulates Melanogenesis in B16F10 Melanoma Cells through the PKA, MAPK, and GSK3β/β-Catenin Signaling Pathways
Next Article in Special Issue
Discovery of a Novel Trifluoromethyl Diazirine Inhibitor of SARS-CoV-2 Mpro
Previous Article in Journal
Cannabidiol as Self-Assembly Inducer for Anticancer Drug-Based Nanoparticles
Previous Article in Special Issue
Medicinal Chemistry of Anti-HIV-1 Latency Chemotherapeutics: Biotargets, Binding Modes and Structure-Activity Relationship Investigation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Expression and Molecular Modification of Chitin Deacetylase from Streptomyces bacillaris

1
Qingdao Key Laboratory of Food Biotechnology, College of Food Science and Engineering, Ocean University of China, Qingdao 266404, China
2
Key Laboratory of Biological Processing of Aquatic Products, China National Light Industry, Qingdao 266404, China
3
Laboratory for Marine Drugs and Bioproducts of Qingdao National Laboratory for Marine Science and Technology, Qingdao 266237, China
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(1), 113; https://doi.org/10.3390/molecules28010113
Submission received: 17 November 2022 / Revised: 19 December 2022 / Accepted: 19 December 2022 / Published: 23 December 2022

Abstract

:
Chitin deacetylase can be used in the green and efficient preparation of chitosan from chitin. Herein, a novel chitin deacetylase SbCDA from Streptomyces bacillaris was heterologously expressed and comprehensively characterized. SbDNA exhibits its highest deacetylation activity at 35 °C and pH 8.0. The enzyme activity is enhanced by Mn2+ and prominently inhibited by Zn2+, SDS, and EDTA. SbCDA showed better deacetylation activity on colloidal chitin, (GlcNAc)5, and (GlcNAc)6 than other forms of the substrate. Molecular modification of SbCDA was conducted based on sequence alignment and homology modeling. A mutant SbCDA63G with higher activity and better temperature stability was obtained. The deacetylation activity of SbCDA63G was increased by 133% compared with the original enzyme, and the optimal reaction temperature increased from 35 to 40 °C. The half-life of SbCDA63G at 40 °C is 15 h, which was 5 h longer than that of the original enzyme. The improved characteristics of the chitin deacetylase SbCDA63G make it a potential candidate to industrially produce chitosan from chitin.

1. Introduction

Chitin (chemically named β-(1,4)-2-acetylamino-2-deoxy-D-glucose) is the second most abundant natural polymer after cellulose and the most abundant renewable nitrogenous material on Earth [1,2]. It is widely found in crustaceans, insect exoskeletons, mollusks, and fungal cell walls [3]. The amount of chitin biosynthesized each year is approximately 100 billion tons, which can be regarded as an inexhaustible biological resource [4].
Chitosan (β-(1,4)-2-amino-2-deoxy-D-glucose) is the deacetylation product of chitin [5]. Due to the presence of amino groups, chitosan has better solubility than chitin [6]. Furthermore, it has good degradability and biocompatibility, and it also has a good hemostatic property, antibacterial property, moisturizing property, formability, and so on [7,8,9,10]. All of the properties mentioned above make chitosan widely used in food processing and preservation, wastewater treatment, medical material, agricultural planting, and many other fields [11,12,13,14].
In industry, chitosan is traditionally prepared by treating chitin with a concentrated alkali [15]. The quality of chitosan is mainly affected by the concentration of the alkali solution, pyrolysis time, and temperature during the reaction [16,17]. The alkali treatment method creates a large amount of environmental pollution, and it is furthermore difficult to control the molecular weight of the product. Consequently, the homogeneity of the produced chitosan is poor, and meeting requirements for the high-quality chitosan in biological and medical fields is difficult [18]. On the contrary, the enzymatic preparation of chitosan from chitin has attracted attention because of its green and controllable reaction process.
Chitin deacetylases (CDA) from the carbohydrate esterase 4 (CE4) family have attracted wide attention because of their ability to remove the acetyl group and produce chitosan or chitooligosaccharide using chitin or chitinosaccharides [19,20,21,22]. However, the studies and research progress regarding CDA are less advanced than those of chitosanase and chitinase. At present, a CDA that has high efficiency and good stability and is suitable for industrial application has not been discovered or developed [23,24,25,26,27,28,29,30,31,32,33]. There is an urgent need to excavate more CDAs and investigate their functional properties with the purpose of furthering the understanding of CDAs and developing CDAs for industrial applications.
In this study, a gene encoding chitin deacetylase was extracted from the genome of Streptomyces bacillaris and heterologously expressed in Bacillus subtilis WB800. The properties of chitin deacetylase were comprehensively studied, and the structure of chitin deacetylase was modified using single-point mutation and saturation mutagenesis. A modified enzyme with improved catalytic properties was obtained. These results provide a good reference for the exploration and molecular modification of new chitin deacetylases.

2. Results

2.1. Bioinformatics Analysis of SbCDA

The open reading frame (ORF) of SbCDA (the amino acid sequences have been provided in Appendix B) consists of 867 nucleotides, encoding a protein of 289 amino acid residues. The molecular mass and isoelectric point of SbCDA were predicted to be 31.37 kDa and 8.60, respectively. To analyze the key catalytic sites, we compared the protein sequence of SbCDA with those of other chitin deacetylases (PesCDA (Genbank: A0A1L3THR9) [34], PdaA (Genbank: AFQ56715) [25], ClCDA (Genbank: Q6DWK3) [35], AnCDA (Genbank: XP_659456) [30], PgtCDA (Genbank: XP_003323413) [36], and AnCDA2 (Genbank: ACF22101.1)). There were 21 completely conserved amino acid sites among these CDAs (as shown by the red squares in Figure 1).

2.2. Expression and Purification of SbCDA

Bacillus subtilis is often used as the traditional fermentation host in industrial production owing to its safety, efficient protein secretion system, and short fermentation period [37]. B. subtilis WB800 is a commonly used expression host which is knocked out of eight extracellular protease genes for the purpose of avoiding the hydrolysis of the target protein [38]. The gene encoding the original SbCDA was cloned and successfully expressed in B. subtilis WB800 with a C-terminal His tag, and the recombinant SbCDA was purified using a Ni-NTA Superflow column. The protein concentrations of crude enzyme and purified SbCDA were 2.2 mg/mL and 0.1 mg/mL, respectively. In addition, the specific enzyme activities of crude enzyme and purified SbCDA under standard condition were 17.1 U/mg and 49.6 U/mg, respectively. The purification factor was 2.0. The product was identified using an Agilent 6460 LC/MS, and the results (Figure A1) confirmed the production of glucosamine (GlcN). SDS–PAGE analysis (Figure A2) showed that the obtained bands were consistent with the predicted results, with a molecular weight of approximately 32 kDa.

2.3. Characterization of SbCDA

2.3.1. Effects of Temperature and pH on the Activity and Stability of SbCDA

The effects of temperature on SbCDA were determined, and the results are shown in Figure 2A. The optimal temperature of SbCDA was 35 °C under the reaction conditions of a pH 7.0 Tris-HCl buffer. The residual enzyme activity (Figure 2A) of SbCDA was tested after incubation at different temperatures for 6 h. The results showed that the activity of SbCDA could be maintained at about 85% at 35 °C and above 95% at 30 °C and 45 °C. Overall, the stability of SbCDA was better below 50 °C, with the remaining enzyme activity being above 80% after 6 h. Residue activity of SbCDA was measured every 4 h at 30, 35, and 40 °C; 24 h was the last time point to be analyzed. The results showed that the half-lives of SbCDA (Figure 2B) at 30, 35, and 40 °C were 24 h, 14 h, and 10 h, respectively.
The effects of pH on the enzyme activity of SbCDA were determined (Figure 2C). The optimum pH of SbCDA was 8.0 in the Tris-HCl buffer solution. Due to the incubation temperature of 40 °C, under which the enzyme activity was maintained well, the incubation time was extended when measuring the stability of SbCDA at different pH conditions. After incubation for 12 h, SbCDA maintained more than 70% of its activity at pH values ranging from 7.0 to 9.0 (Figure 2D).

2.3.2. Effects of Chemical Reagents on SbCDA Enzyme Activity

To investigate the effects of different metal ions on SbCDA and the tolerance of SbCDA to protein-deformable agents and to confirm that chitin deacetylase is a metal ion-dependent enzyme, different metal ions (Fe3+, Ca2+, Cu2+, Mg2+, Zn2+, Mn2+, Ni2+, Ba2+, Co2+, K+, and Na+), SDS, and Na2EDTA were added into the reaction system, respectively. The results (Figure 3) showed that the activity of SbCDA was strongly inhibited by Zn2+ and was mildly inhibited by Ca2+, Cu2+, Mg2+, Ba2+, and Na2+. The inhibition became stronger at high concentrations. Fe3+, Ni2+, and Co2+ slightly improved the activity of SbCDA, while Mn2+ strongly improved the activity of SbCDA. K+ inhibited SbCDA activity at 1 mM but enhanced the activity at 10 mM. SDS and Na2EDTA also strongly inhibited the activity of SbCDA.

2.4. Substrate Preference of SbCDA

By selecting chitin oligosaccharides with different degrees of polymerization and chitins with different solubilities as reaction substrates, the enzyme activity of SbCDA with different substrates was determined. As shown in Figure 4, SbCDA showed activity against colloidal chitin and soluble chitin oligosaccharides and had no effect on insoluble powder chitin. Compared with carboxymethyl chitin, GlcNAc, and (GlcNAc)2–4, SbCDA clearly prefers colloidal chitin, (GlcNAc)5, and (GlcNAc)6.

2.5. Comparison of Enzyme Activity of Mutants and Analysis of Homologous CDAs

By sequence alignment, we selected 19 amino acids (the blue dots in Figure 1) in the conserved region and mutated them into alanine. The enzyme activity of these 19 mutants was studied (Figure 5). The mutants SbCDA62, SbCDA63, SbCDA77, SbCDA79, SbCDA87, SbCDA88, SbCDA156, and SbCDA208 showed relatively obvious changes in enzyme activity. Then, to explore the role of the aforementioned amino acids in the structure, we established a homology model of SbCDA using eight crystal structures as templates (PDB ID: 5LFZ [39], 6H8L [40], 2C1G [41], 5NC6 [42], 7AX7 (https://www1.rcsb.org/structure/7AX7, accessed on 16 November 2022), 1NY1 (https://www1.rcsb.org/structure/1NY1, accessed on 16 November 2022), 5O6Y (https://www1.rcsb.org/structure/5O6Y, accessed on 16 November 2022), and 4L1G [43]). The homology model (Figure 6) indicated that the His116 residues and His120 residues can form coordination bonds with Zn2+ and that Asp65 and His210 are the general base residue and the general acid residue for catalysis, respectively [21]. The three-dimensional protein structure diagram (Figure 7) shows that Leu62, Thr63, Leu79, Thr87, Phe88, and Leu208 of SbCDA are around the catalytic site of the metal triplet. Therefore, we selected SbCDA62, SbCDA63, SbCDA79, SbCDA87, SbCDA88, and SbCDA208 for saturation mutation to search for mutants with higher enzyme activity.

2.6. SbCDA Saturation Mutations at Six Amino Acid Sites

Based on the results of single-point mutation and structural analysis, saturation mutations were prepared on six amino acids at positions 62, 63, 79, 87, 88, and 208. The mutants were cultured, and the mutant proteins were collected under the same conditions. In order to comprehensively evaluate the enzyme activity and expression level, the activity of each mutant was determined by adding the same amount of supernatant (400 μL). As shown in Figure 8, mutant SbCDA63G with Thr63 mutated to Gly, mutant SbCDA79H with Leu79 mutated to His, and mutant SbCDA87R with Thr87 mutated to Arg had relatively good enzyme activity. These three mutant strains were cultured, and the crude enzymes were collected. However, only mutant SbCDA63G could be purified by Ni column chromatography. The protein concentrations of crude SbCDA63G and purified SbCDA63G were 1.9 mg/mL and 0.2 mg/mL, respectively. In addition, the specific enzyme activities of crude enzyme and purified SbCDA were 35.9 U/mg and 115.9 U/mg, respectively. The purification factor was 3.2. SDS–PAGE analysis (Figure A3) showed that the obtained protein band was consistent with the predicted result. The mutant SbCDA63G protein had a molecular weight of approximately 32 kDa.

2.7. Characterization of Mutant SbCDA63G

When it comes to the optimum temperature, SbCDA63G displayed its highest activity at 40 °C. When it comes to the stability of the enzyme at different temperature, SbCDA63G showed more than 75% residual activity from 25 °C to 65 °C after being incubated for 6 h (Figure 9A). The half-lives of SbCDA63G at 35, 40, and 45 °C were 12, 15, and 24 h, respectively (Figure 9B). The optimal pH of SbCDA63G was 8.0. More than 75% of the activity of SbCDA63G was retained after incubation for 12 h in a pH range from 7.0 to 9.0 (Figure 9D). SbCDA63G has the greatest stability in a Tris-HCl buffer solution at pH 7.0.

3. Discussion

In this study, SbCDA from S. bacillaris was obtained and investigated. The optimal temperature of SbCDA was 35 °C under the reaction conditions of pH 7.0 Tris-HCl buffer (Figure 2A). In general, the optimum temperatures of CDAs from fungi (generally 50–60 °C) are higher than those of CDAs from bacteria [28,30,44,45,46,47,48,49]. Theoretically, the higher the temperature, the lower the stability of the enzyme, but in this study, the stability became higher at 45 °C (Figure 2A), probably because of the activation effect of temperature on the enzyme, but this phenomenon has not been found in other chitin deacetylases [30,33,34,35,36,39,40,41,42,50,51]. In addition, SbCDA exhibited below 50% residual activity after incubation at 30, 35, and 40 °C for 24 h, indicating that the stability of SbCDA is not very good. Therefore, one of the purposes of mutation is to improve the stability of SbCDA. The optimum pH of 8.0 for SbCDA is similar to those of CDAs derived from Bacillus licheniformis, Nitratireductor aquimarinus MCDA3-3, and Vibrio cholerae [45,52,53]. The high activity and good stability of SbCDA in Tris-HCl buffer from pH 7.0–9.0 further support the idea that SbCDA is an alkaline chitin deacetylase [54]. Activation by Mn2+ and inhibition by Na2EDTA indicate that SbCDA is metal ion-dependent. In fact, the relationships between metal ions and the activity of almost all CDAs have been studied. For instance, Cd2+, Co2+, and EDTA ions inhibited the activity of MeCDA from M. esteraromaticum while K+, Li+, and Sr2+ ions strongly enhanced the enzyme’s activity [54]. ClCDA from Colletotrichum lindemuthianum was inhibited by Zn2+, Mn2+, and Cu2+, whereas it was not inhibited by Na+, K+, Li+, Mg2+, or Ca2+. Co2+ could improve the activity of ClCDA [55]. We cannot say for certain which metal ions have an inhibitory effect on the activity of CDAs and which metal ions could promote their activity, but it is worth making sure that most CDAs are metal ion-dependent and that EDTA can inhibit the activity of CDAs. In terms of substrate preference, SbCDA and most reported CDA enzymes are mainly active on soluble chitins, such as colloidal chitin, and soluble chitosaccharides, but they have no activity against natural chitin [56]. However, MCDA02 from Microbacterium esteraromaticum and MCDA3-3 from Nitratireductor aquimarinus showed deacetylase activity toward α-chitin [54,56]. Most CDAs have no activity toward GlcNAc, while SbCDA in our study has activity toward GlcNAc. If SbCDA were used in combination with other enzymes that have different substrate preferences, different substrates could be utilized more efficiently, and a wide variety of chitosan oligosaccharides which have various physiological activities could be produced.
The enzymatic activity and stability of SbCDA did not meet our expectations, and we modified CDA by single-point mutation and saturation mutation based on multiple sequence alignment to obtain a more efficient and stable CDA. The result of sequence alignment showed that there were 21 completely conserved amino acid sites among CDAs participating in the comparison (as shown by the red squares in Figure 1). According to previous research, CDAs operate by metal-assisted acid/base catalysis [57]. Asp65 and His210 (as shown by purple triangles in Figure 1) are the general base residue and the general acid residue for catalysis, respectively [21]. The other 19 completely conserved sites of CDAs are marked by blue dots. Previous studies have suggested conserved motifs that affect enzyme activity; however, this consensus can be improved [21]. Therefore, we speculated that our 19 highly conserved sites might influence the activity of SbCDA, and we mutated these 19 amino acids (the blue dots in Figure 1) in the conserved region into alanine, which is a chiral amino acid with the shortest side chain [58]. Then, we studied the enzyme activity of these 19 mutants.
Among the 19 mutants, SbCDA62, SbCDA63, SbCDA77, SbCDA79, SbCDA87, SbCDA88, SbCDA156, and SbCDA208 showed the greatest alterations in enzyme activity, indicating that the change of amino acids at these positions is promising to improve the enzymatic activity of the enzyme. Meanwhile, the results of homology modeling illustrated that Leu62, Thr63, Leu79, Thr87, Phe88, and Leu208 were near the catalytic site of the metallic triad. Combining these two aspects, six amino acid sites (positions 62, 63, 79, 87, 88, and 208) underwent saturation mutagenesis. Among all mutants, SbCDA63G, SbCDA79H, and SbCDA87R had relatively good enzyme activity. However, only SbCDA63G could be purified. Therefore, the next inquiry of enzymatic properties was conducted using the mutant SbCDA63G.
SbCDA63G has higher activity and better stability compared with SbCDA (Figure 9). The specific enzyme activity of SbCDA63G was 115.9 U/mg, while that of SbCDA was 49.6 U/mg, indicating that the activity of SbCDA63G increased by approximately 133%. The increase in activity may be due to the increased flexibility of the main chain and less entropy loss during folding after mutation to Gly [59]. SbCDA63G had better thermostability than SbCDA, suggesting that the mutant is more suitable for industrial applications than SbCDA [60]. A possible explanation for the increase in thermal stability is that the interaction between Gly and the nearby charged amino acids (D, E, K, and R) is more beneficial to the stability of the protein [61]. The mutation did not affect the optimal pH of the enzyme, which was 8.0 for both SbCDA63G and SbCDA (Figure 9C and Figure 2C). SbCDA63G has its greatest stability in Tris-HCl buffer solution at pH 7.0, compared to pH 8.0 for SbCDA.
Our study indicates that Thr63 has a great effect on enzyme activity. The mutation of Thr to Gly could improve the activity and stability of SbCDA, making it more suitable for industrial production.

4. Materials and Methods

4.1. Materials

A One Step Cloning Kit and Taq polymerase were purchased from Vazyme (Nanjing, China). High-fidelity DNA polymerase was obtained from Toyobo (Osaka, Japan). K-ACET was purchased from Megazyme (Bray, Ireland). A Bacteria DNA Kit and a Plasmid Extraction Kit were purchased from Tiangen (Beijing, China). A Gel Extraction Kit was purchased from Omega (Guangzhou, China). Chitin was purchased from Sinopharm Group (Shanghai, China). Ni-NTA Superflow was obtained from Qiagen (Duesseldorf, Germany). N-acetyl glucosamine (GlcNAc), diacetylchitobiose ((GlcNAc)2), triacetylchitotriose ((GlcNAc)3), tetraacetylchitotetraose ((GlcNAc)4), acetylchitopentaose ((GlcNAc)5), and acetylchitohexaose ((GlcNAc)6) were purchased from Qingdao Bozhihuili Biotechnology (Qingdao, China). Other chemical reagents used in this study were of analytical grade unless specifically indicated. Colloidal chitin and carboxymethyl chitin were prepared according to previously reported methods [50,51].

4.2. Strains and Culture Conditions

Streptomyces bacillaris was activated in tryptic soy broth (TSB) medium containing 1.5% (w/v) tryptone, 0.5% (w/v) soya peptone, and 1% (w/v) NaCl. DH5α chemically competent cells were cultivated at 37 °C in Luria–Bertani (LB) medium composed of 0.5% (w/v) yeast extract, 1% (w/v) tryptone, and 1% (w/v) NaCl with 0.05 g/L kanamycin when needed. Bacillus subtilis WB800 was grown in LB medium at 37 °C. Thallus growth medium (GM) used for culturing B. subtilis WB800 contained 0.5% (w/v) yeast extract, 1% (w/v) tryptone, 1% (w/v) NaCl, and 9.1% (w/v) sorbitol. Electrotransfer medium (ETM) for the preparation of competent B. subtilis cells contained 9.1% (w/v) sorbitol, 9.1% (w/v) mannitol, and 10% (v/v) glycerol. Thallus recovery medium (RM) for the resuscitation of transformed cells contained 0.5% (w/v) yeast extract, 1% (w/v) tryptone, 1% (w/v) NaCl, 9.1% (w/v) sorbitol, and 6.9% (w/v) mannitol.

4.3. Bioinformatics Analysis, Homology Modeling, and Molecular Docking of CDAs

ExPASy ProtParam (https://web.expasy.org/protparam/, accessed on 16 November 2022) was used to calculate the isoelectric point and molecular weight. Both ESPript and ClustalX 2.1 were used to compare the amino acid sequences of chitin deacetylases (PesCDA (Genbank: A0A1L3THR9) [34], PdaA (Genbank: AFQ56715) [25], ClCDA (Genbank: Q6DWK3) [35], AnCDA (Genbank: XP_659456) [30], PgtCDA (Genbank: XP_003323413) [39], and AnCDA2 (Genbank: ACF22101.1)) with SbCDA. The MODELLER software (version 9.23) was used to establish the homology model of the chitin deacetylase SbCDA using eight crystal structures (PDB ID: 5LFZ [39], 6H8L [40], 2C1G [41], 5NC6 [42], 7AX7 (https://www1.rcsb.org/structure/7AX7, accessed on 16 November 2022), 1NY1 (https://www1.rcsb.org/structure/1NY1, accessed on 16 November 2022), 5O6Y (https://www1.rcsb.org/structure/5O6Y, accessed on 16 November 2022), and 4L1G [43]) as templates. The geometric conformation of the protein model was observed using PyMOL (Version 2.4.1). SAVES (http://services.mbi.ucla.edu/SAVES/, accessed on 16 November 2022), PROCHECK ERRAT, and VERIFY3D were used for further evaluation of the geometric conformation of the protein model. The molecular structure of GlcNAc was obtained from the ZINC database (http://zinc.docking.org/, accessed on 16 November 2022). The AutoDock (version 4.2.6) software (http://mgltools.scripps.edu/, accessed on 16 November 2022) was used for molecular docking. Finally, the docking site conforming to the catalytic conditions was selected as the initial conformation for analysis.

4.4. Construction of the SbCDA Expression Plasmid and Site-Directed Mutation

Genomic DNA was extracted from S. bacillaris using the Bacteria DNA Kit. Primer synthesis for the amplification of the SbCDA gene, construction of mutant plasmids, and linearization of the p43NMK vector were conducted at Sangon Biotech Co., Ltd. (Shanghai, China). PCR products were purified and then ligated into the pP43NMK vector, which contains a 6×His tag, according to the instructions of the One Step Cloning Kit manufacturer. The recombinant and mutant plasmids were transformed into DH5α chemically competent cells for propagation of the plasmids. The Plasmid Extraction Kit was used to recover the plasmids. Then, the recombinant and mutant plasmids were electrotransformed into B. subtilis WB800 chemically competent cells to obtain transformants and mutant strains, respectively.

4.5. Expression and Purification of SbCDA and the Mutant Enzymes

To express the target enzymes, the transformants and mutant strains were grown in LB medium with shaking (220 rpm) for 20 h at 37 °C. After culture, the supernatant was obtained by centrifugation at 8000 rpm for 15 min at 4 °C to collect the crude enzyme solution. The activity of the crude enzyme solution was measured, and the mutant with an obvious enzyme activity change was selected for saturation mutation. Then, the mutant with the highest enzyme activity was selected to evaluate its properties. The enzyme was purified by a Ni-NTA Superflow column eluted with a gradient of imidazole (10, 50, 80, 100, 200, and 500 mM) in 50 mM Tris-HCl buffer (pH 8.0) and 500 mM NaCl. Finally, the purified protein was analyzed by SDS–PAGE. The content of protein was measured by Bradford assay [62]. The purified enzyme was used for further enzyme activity assays and biochemical characterization.

4.6. Chitin Deacetylase Activity Assay

Within all activity and stability tests, the enzyme activity assay was performed as previously described using the coupled enzymatic method with some modifications [36]. For the assay, 400 μL GlcNAc (10 g/L) was added to 400 μL Tris-HCl buffer solution (0.01 mol/L, pH 7.0), followed by 100 μL of enzyme solution, and the reaction system was incubated at 40 °C for 6 h. The activity of chitin deacetylase was characterized by measuring the content of acetic acid produced using the K-ACET acetic acid determination kit (Megazyme, Bray, Ireland).
One unit of enzymatic activity (U) was defined as the amount of enzyme (mg) required to produce 1 μmol of acetic acid per minute under the above conditions.
In order to verify the deacetylation products of chitin deacetylases, the hydrolysates were analyzed by mass spectrometry using an Agilent 6460 LC/MS.

4.7. Biochemical Characterization of SbCDA and the Mutant Enzymes

The impact of temperature on SbCDA activity under standard conditions was studied from 25 to 65 °C, with a temperature interval of 5 °C. The same amount of enzyme solution was kept at the above temperature for 6 h to determine the stability. For the thermal stability assay, aliquots of enzyme were placed at 30, 35, and 40 °C for 0, 4, 8, 12, 16, 20, and 24 h. The residual enzyme activity was measured under standard conditions to determine the thermal stability. The enzyme half-life is calculated as the time it takes for the enzyme activity to drop to half of the initial enzyme activity.
Under standard conditions, the effect of pH buffers was examined in citric acid buffer (pH 3.0–6.0), phosphoric acid buffer (pH 6.0–7.0), Tris-HCl buffer (pH 7.0–9.0), and glycine sodium hydroxide buffer (pH 9.0–10.0). The concentration of all buffers was 10 mM. To investigate the effects of pH on the stability of SbCDA, the enzyme was incubated in citric acid buffer (pH 3.0–6.0), phosphoric acid buffer (pH 6.0–7.0), Tris-HCl buffer (pH 7.0–9.0), and glycine sodium hydroxide buffer (pH 9.0–10.0) for 12 h, respectively. The residual activity was measured according to the standard method.

4.8. Effects of Metal Ions and Chemical Reagents on SbCDA Activity

To examine the effects of metal ions (Fe3+, Ca2+, Cu2+, Mg2+, Zn2+, Mn2+, Ni2+, Ba2+, Co2+, K+, and Na+) and a series of chemicals (disodium ethylenediaminetetraacetate (Na2EDTA) and sodium dodecyl sulfate (SDS)) on the activity of SbCDA, they were added to the reaction mixture at final concentrations of 1 and 10 mM. The metal ions were each added to an enzyme solution, which was incubated for 1 h at 37 °C. Then, to measure the relative activity of SbCDA, the reaction system was incubated at 40 °C for 6 h in Tris-HCl buffer (10 mM, pH = 7.0). The control group did not have metal ions or chemical reagents added under the same reaction system and reaction conditions as the experimental group.

4.9. Substrate Preference of SbCDA

Under standard conditions, chitin deacetylase was reacted with chitin, colloidal chitin, carboxymethyl chitin, GlcNAc, and (GlcNAc)2 to (GlcNAc)6 at a concentration of 10 mg/mL. The activity of chitin deacetylase on different substrates was determined.

5. Conclusions

In summary, a novel chitin deacetylase SbCDA from S. bacillaris was expressed, purified, characterized, and molecularly modified. The phenomenon that Mn2+ can strongly improve the activity of SbCDA needs to be further investigated to elucidate the activation mechanism of Mn2+. Different substrate preferences of SbCDA from previously studied enzymes provide biological tools with potential for the production of multiple chitosan types. The activity of the mutant enzyme SbCDA63G was 133% higher than that of SbCDA. We speculate that the reason for this phenomenon is that Gly has low spatial bit resistance. When Thr was mutated to Gly at position 63, the entropy loss during folding decreased, and the flexibility of the main chain increased, which made it easier for the active center of the enzyme to deform and bind with the substrate. Moreover, the interaction between Gly and the nearby charged amino acids (D, E, K, and R) is more beneficial to the stability of the protein, and thus the optimal temperature and thermal stability of SbCDA63G were also improved, making it more suitable for industrial production. Improved enzyme activity and stability would allow the mutant SbCDA63G to achieve the same deacetylation as the original enzyme with less dosage. Our study demonstrates that mutation to Gly improves protein stability, providing a useful method for improving the properties of a chitin deacetylase and a potential tool for preparing chitosan or chitosan oligosaccharides from chitin.

Author Contributions

Conceptualization, L.Y. and X.M.; data curation, L.Y. and Q.W.; funding acquisition, J.S. and X.M.; investigation, L.Y. and Q.W.; methodology, J.S.; supervision, J.S.; visualization, L.Y., Q.W., and J.S.; writing—original draft, L.Y. and X.M.; writing—review and editing, J.S. and X.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China, grant number U2106228; Qingdao Shinan District Science and Technology Plan Project 2022-3-010-SW; the China Agriculture Research System of MOF and MARA, grant number CARS-48; and the Taishan Scholar Project of Shandong Province, grant number tsqn201812020.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not available.

Appendix A

Figure A1. MS analysis of products of GlcNAc.
Figure A1. MS analysis of products of GlcNAc.
Molecules 28 00113 g0a1
The mass–charge ratio peaks of 162.1 and 180.1 in the mass spectrogram result are fragment ions of [GlcN+H]+ and [GlcN-H2O+H]+, respectively, which confirms the production of glucosamine (GlcN).
Figure A2. SDS-PAGE analysis of purified SbCDA. M, molecular weight marker; 1, purified SbCDA.
Figure A2. SDS-PAGE analysis of purified SbCDA. M, molecular weight marker; 1, purified SbCDA.
Molecules 28 00113 g0a2
Figure A3. SDS-PAGE analysis of purified SbCDA63G. Lane M: protein molecular mass marker; Lane 1: purified SbCDA63G.
Figure A3. SDS-PAGE analysis of purified SbCDA63G. Lane M: protein molecular mass marker; Lane 1: purified SbCDA63G.
Molecules 28 00113 g0a3

Appendix B

The sequence of SbCDA is as follows:
  • >SbCDAMPKKMSVLGGGLAAALVATLTLALTGCSMETTAPASARRDAAPDAKGSFGRADCRKAKCIALTFDAGPGKDTAELLDILKEKKVSATFFLLGRNHVLKHPDTVRRIQDEGHEVANHTWTHKILTDEKPEEIRAELEKTQEAIEKITGKRPRLMRPPQGRTDDQVSGISKELGLSQVLWSATAKDYSTNDSALITQRILDQASRDGIILLHDIYKGTVPAVPGIIDALQKDGYTFVTVPELMAPAVPEPGTIYRP

References

  1. Younes, I.; Rinaudo, M. Chitin and Chitosan Preparation from Marine Sources. Structure, Properties and Applications. Mar. Drugs 2015, 13, 1133–1174. [Google Scholar] [CrossRef] [Green Version]
  2. Ma, X.; Gözaydın, G.; Yang, H.; Ning, W.; Han, X.; Poon, N.Y.; Liang, H.; Yan, N.; Zhou, K. Upcycling chitin-containing waste into organonitrogen chemicals via an integrated process. Proc. Natl. Acad. Sci. USA 2020, 117, 7719–7728. [Google Scholar] [CrossRef]
  3. Tharanathan, R.N.; Kittur, F.S. Chitin—The Undisputed Biomolecule of Great Potential. Crit. Rev. Food Sci. Nutr. 2003, 43, 61–87. [Google Scholar] [CrossRef]
  4. Yan, N.; Chen, X. Sustainability: Don’t waste seafood waste. Nature 2015, 524, 155–157. [Google Scholar] [CrossRef] [Green Version]
  5. Liaqat, F.; Eltem, R. Chitooligosaccharides and their biological activities: A comprehensive review. Carbohydr. Polym. 2018, 184, 243–259. [Google Scholar] [CrossRef]
  6. de Assis, C.F.; Araújo, N.K.; Pagnoncelli, M.G.B.; da Silva Pedrini, M.R.; de Macedo, G.R.; dos Santos, E.S. Chitooligosaccharides enzymatic production by Metarhizium anisopliae. Bioprocess. Biosyst. Eng. 2010, 33, 893–899. [Google Scholar] [CrossRef]
  7. Nawrotek, K.; Tylman, M.; Decherchi, P.; Marqueste, T.; Rudnicka, K.; Gatkowska, J.; Wieczorek, M. Assessment of degradation and biocompatibility of electrodeposited chitosan and chitosan–carbon nanotube tubular implants. J. Biomed. Mater. Res. Part A 2016, 104, 2701–2711. [Google Scholar] [CrossRef]
  8. Hu, Z.; Lu, S.; Cheng, Y.; Kong, S.; Li, S.; Li, C.; Yang, L. Investigation of the Effects of Molecular Parameters on the Hemostatic Properties of Chitosan. Molecules 2018, 23, 3147. [Google Scholar] [CrossRef] [Green Version]
  9. Sahariah, P.; Másson, M. Antimicrobial Chitosan and Chitosan Derivatives: A Review of the Structure–Activity Relationship. Biomacromolecules 2017, 18, 3846–3868. [Google Scholar] [CrossRef]
  10. Chaiwong, N.; Leelapornpisid, P.; Jantanasakulwong, K.; Rachtanapun, P.; Seesuriyachan, P.; Sakdatorn, V.; Leksawasdi, N.; Phimolsiripol, Y. Antioxidant and Moisturizing Properties of Carboxymethyl Chitosan with Different Molecular Weights. Polymers 2020, 12, 1445. [Google Scholar] [CrossRef]
  11. Lai, W.-F.; Zhao, S.; Chiou, J. Antibacterial and clusteroluminogenic hypromellose-graft-chitosan-based polyelectrolyte complex films with high functional flexibility for food packaging. Carbohydr. Polym. 2021, 271, 118447. [Google Scholar] [CrossRef]
  12. Bhatnagar, A.; Sillanpää, M. Applications of chitin- and chitosan-derivatives for the detoxification of water and wastewater—A short review. Adv. Colloid Interface Sci. 2009, 152, 26–38. [Google Scholar] [CrossRef]
  13. Khan, A.; Wang, B.; Ni, Y. Chitosan-Nanocellulose Composites for Regenerative Medicine Applications. Curr. Med. Chem. 2020, 27, 4584–4592. [Google Scholar] [CrossRef]
  14. Shahrajabian, M.H.; Chaski, C.; Polyzos, N.; Tzortzakis, N.; Petropoulos, S.A. Sustainable Agriculture Systems in Vegetable Production Using Chitin and Chitosan as Plant Biostimulants. Biomolecules 2021, 11, 819. [Google Scholar] [CrossRef]
  15. No, H.K.; Meyers, S.P. Preparation and Characterization of Chitin and Chitosan—A Review. J. Aquat. Food Prod. Technol. 1995, 4, 27–52. [Google Scholar] [CrossRef]
  16. Kumari, S.; Rath, P.; Sri Hari Kumar, A.; Tiwari, T.N. Extraction and characterization of chitin and chitosan from fishery waste by chemical method. Environ. Technol. Innov. 2015, 3, 77–85. [Google Scholar] [CrossRef]
  17. Kou, S.; Peters, L.M.; Mucalo, M.R. Chitosan: A review of sources and preparation methods. Int. J. Biol. Macromol. 2021, 169, 85–94. [Google Scholar] [CrossRef]
  18. Chen, X.; Yang, H.; Zhong, Z.; Yan, N. Base-catalysed, one-step mechanochemical conversion of chitin and shrimp shells into low molecular weight chitosan. Green Chem. 2017, 19, 2783–2792. [Google Scholar] [CrossRef] [Green Version]
  19. Cantarel, B.L.; Coutinho, P.M.; Rancurel, C.; Bernard, T.; Lombard, V.; Henrissat, B. The Carbohydrate-Active EnZymes database (CAZy): An expert resource for Glycogenomics. Nucleic Acids Res. 2009, 37, D233–D238. [Google Scholar] [CrossRef]
  20. John, M.; Röhrig, H.; Schmidt, J.; Wieneke, U.; Schell, J. Rhizobium NodB protein involved in nodulation signal synthesis is a chitooligosaccharide deacetylase. Proc. Natl. Acad. Sci. USA 1993, 90, 625–629. [Google Scholar] [CrossRef]
  21. Aragunde, H.; Biarnés, X.; Planas, A. Substrate Recognition and Specificity of Chitin Deacetylases and Related Family 4 Carbohydrate Esterases. Int. J. Mol. Sci. 2018, 19, 412. [Google Scholar] [CrossRef] [Green Version]
  22. Zhao, Y.; Park, R.-D.; Muzzarelli, R.A.A. Chitin Deacetylases: Properties and Applications. Mar. Drugs 2010, 8, 24–46. [Google Scholar] [CrossRef]
  23. ElMekawy, A.; Hegab, M.H.; El-Baz, A.; Hudson, M.S. Kinetic Properties and Role of Bacterial Chitin Deacetylase in the Bioconversion of Chitin to Chitosan. Recent Pat. Biotechnol. 2013, 7, 234–241. [Google Scholar] [CrossRef]
  24. Raval, R.; Simsa, R.; Raval, K. Expression studies of Bacillus licheniformis chitin deacetylase in E. coli Rosetta cells. Int. J. Biol. Macromol. 2017, 104, 1692–1696. [Google Scholar] [CrossRef]
  25. Blair, D.E.; van Aalten, D.M.F. Structures of Bacillus subtilis PdaA, a family 4 carbohydrate esterase, and a complex with N-acetyl-glucosamine. FEBS Lett. 2004, 570, 13–19. [Google Scholar] [CrossRef] [Green Version]
  26. Andreou, A.; Giastas, P.; Christoforides, E.; Eliopoulos, E.E. Structural and Evolutionary Insights within the Polysaccharide Deacetylase Gene Family of Bacillus anthracis and Bacillus cereus. Genes 2018, 9, 386. [Google Scholar] [CrossRef] [Green Version]
  27. Urch, J.E.; Hurtado-Guerrero, R.; Brosson, D.; Liu, Z.; Eijsink, V.G.H.; Texier, C.; van Aalten, D.M.F. Structural and functional characterization of a putative polysaccharide deacetylase of the human parasite Encephalitozoon cuniculi. Protein Sci. 2009, 18, 1197–1209. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Karthik, N.; Binod, P.; Pandey, A. SSF production, purification and characterization of chitin deacetylase from Aspergillus flavus. Biocatal. Biotransform. 2018, 36, 296–306. [Google Scholar] [CrossRef]
  29. Xie, M.; Zhao, X.; Lü, Y.; Jin, C. Chitin deacetylases Cod4 and Cod7 are involved in polar growth of Aspergillus fumigatus. MicrobiologyOpen 2020, 9, e00943. [Google Scholar] [CrossRef] [Green Version]
  30. Liu, Z.; Gay, L.M.; Tuveng, T.R.; Agger, J.W.; Westereng, B.; Mathiesen, G.; Horn, S.J.; Vaaje-Kolstad, G.; van Aalten, D.M.F.; Eijsink, V.G.H. Structure and function of a broad-specificity chitin deacetylase from Aspergillus nidulans FGSC A4. Sci. Rep. 2017, 7, 1746. [Google Scholar] [CrossRef]
  31. Liu, L.; Zhou, Y.; Qu, M.; Qiu, Y.; Guo, X.; Zhang, Y.; Liu, T.; Yang, J.; Yang, Q. Structural and biochemical insights into the catalytic mechanisms of two insect chitin deacetylases of the carbohydrate esterase 4 family. J. Biol. Chem. 2019, 294, 5774–5783. [Google Scholar] [CrossRef] [Green Version]
  32. Sarmiento, K.P.; Panes, V.A.; Santos, M.D. Molecular cloning and expression of chitin deacetylase 1 gene from the gills of Penaeus monodon (black tiger shrimp). Fish Shellfish Immunol. 2016, 55, 484–489. [Google Scholar] [CrossRef]
  33. Fan, B.; Li, Y.; Huang, Y.; Zhang, M.; Liu, Z.; Fan, W.; Zhao, Y. Cloning and expression of chitin deacetylase 1 from Macrobrachium nipponense, and the effects of dietary protein on growth, body composition and digestive enzymes. Aquacult. Nutr. 2018, 24, 1664–1678. [Google Scholar] [CrossRef]
  34. Cord-Landwehr, S.; Melcher, R.L.J.; Kolkenbrock, S.; Moerschbacher, B.M. A chitin deacetylase from the endophytic fungus Pestalotiopsis sp. efficiently inactivates the elicitor activity of chitin oligomers in rice cells. Sci. Rep. 2016, 6, 38018. [Google Scholar] [CrossRef]
  35. Tokuyasu, K.; Ohnishi-Kameyama, M.; Hayashi, K.; Mori, Y. Cloning and expression of chitin deacetylase gene from a deuteromycete, Colletotrichum lindemuthianum. J. Biosci. Bioeng. 1999, 87, 418–423. [Google Scholar] [CrossRef]
  36. Naqvi, S.; Cord-Landwehr, S.; Singh, R.; Bernard, F.; Kolkenbrock, S.; El Gueddari Nour, E.; Moerschbacher Bruno, M. A Recombinant Fungal Chitin Deacetylase Produces Fully Defined Chitosan Oligomers with Novel Patterns of Acetylation. Appl. Environ. Microbiol. 2016, 82, 6645–6655. [Google Scholar] [CrossRef] [Green Version]
  37. Wang, G.; Guo, Z.; Zhang, X.; Wu, H.; Bai, X.; Zhang, H.; Hu, R.; Han, S.; Pang, Y.; Gao, Z.a.; et al. Heterologous expression of pediocin/papA in Bacillus subtilis. Microb. Cell Fact. 2022, 21, 104. [Google Scholar] [CrossRef]
  38. Murashima, K.; Chen, C.-L.; Kosugi, A.; Tamaru, Y.; Doi Roy, H.; Wong, S.-L. Heterologous Production of Clostridium cellulovorans engB, Using Protease-Deficient Bacillus subtilis, and Preparation of Active Recombinant Cellulosomes. J. Bacteriol. 2002, 184, 76–81. [Google Scholar] [CrossRef] [Green Version]
  39. Tuveng, T.R.; Rothweiler, U.; Udatha, G.; Vaaje-Kolstad, G.; Smalas, A.; Eijsink, V.G.H. Structure and function of a CE4 deacetylase isolated from a marine environment. PLoS ONE 2017, 12, 0187544. [Google Scholar] [CrossRef] [Green Version]
  40. Grifoll-Romero, L.; Sainz-Polo, M.A.; Albesa-Jové, D.; Guerin, M.E.; Biarnés, X.; Planas, A. Structure-function relationships underlying the dual N-acetylmuramic and N-acetylglucosamine specificities of the bacterial peptidoglycan deacetylase PdaC. J. Biol. Chem. 2019, 294, 19066–19080. [Google Scholar] [CrossRef]
  41. Blair, D.E.; Schüttelkopf, A.W.; MacRae, J.I.; van Aalten, D.M.F. Structure and metal-dependent mechanism of peptidoglycan deacetylase, a streptococcal virulence factor. Proc. Natl. Acad. Sci. USA 2005, 102, 15429–15434. [Google Scholar] [CrossRef] [Green Version]
  42. Giastas, P.; Andreou, A.; Papakyriakou, A.; Koutsioulis, D.; Balomenou, S.; Tzartos, S.J.; Bouriotis, V.; Eliopoulos, E.E. Structures of the Peptidoglycan N-Acetylglucosamine Deacetylase Bc1974 and Its Complexes with Zinc Metalloenzyme Inhibitors. Biochemistry 2018, 57, 753–763. [Google Scholar] [CrossRef]
  43. Fadouloglou, V.E.; Balomenou, S.; Aivaliotis, M.; Kotsifaki, D.; Arnaouteli, S.; Tomatsidou, A.; Efstathiou, G.; Kountourakis, N.; Miliara, S.; Griniezaki, M.; et al. Unusual α-Carbon Hydroxylation of Proline Promotes Active-Site Maturation. J. Am. Chem. Soc. 2017, 139, 5330–5337. [Google Scholar] [CrossRef]
  44. Wang, Y.; Song, J.-Z.; Yang, Q.; Liu, Z.-H.; Huang, X.-M.; Chen, Y. Cloning of a Heat-Stable Chitin Deacetylase Gene from Aspergillus nidulans and its Functional Expression in Escherichia coli. Appl. Biochem. Biotechnol. 2010, 162, 843–854. [Google Scholar] [CrossRef]
  45. Alfonso, C.; Nuero, O.M.; Santamaría, F.; Reyes, F. Purification of a heat-stable chitin deacetylase from Aspergillus nidulans and its role in cell wall degradation. Curr. Microbiol. 1995, 30, 49–54. [Google Scholar] [CrossRef]
  46. Blair, D.E.; Hekmat, O.; Schüttelkopf, A.W.; Shrestha, B.; Tokuyasu, K.; Withers, S.G.; van Aalten, D.M.F. Structure and Mechanism of Chitin Deacetylase from the Fungal Pathogen Colletotrichum lindemuthianum. Biochemistry 2006, 45, 9416–9426. [Google Scholar] [CrossRef] [Green Version]
  47. Yamada, M.; Kurano, M.; Inatomi, S.; Taguchi, G.; Okazaki, M.; Shimosaka, M. Isolation and characterization of a gene coding for chitin deacetylase specifically expressed during fruiting body development in the basidiomycete Flammulina velutipes and its expression in the yeast Pichia pastoris. FEMS Microbiol. Lett. 2008, 289, 130–137. [Google Scholar] [CrossRef] [Green Version]
  48. Zhao, Y.; Kim, Y.J.; Oh, K.T.; Nguyen, V.N.; Park, R.D. Production and Characterization of Extracellular Chitin Deacetylase from Absidia corymbifera DY-9. J. Korean Soc. Appl. Biol. Chem. 2010, 53, 119–126. [Google Scholar] [CrossRef]
  49. Kafetzopoulos, D.; Martinou, A.; Bouriotis, V. Bioconversion of Chitin to Chitosan—Purification and Characterization of Chitin Deacetylase from Mucor-Rouxii. Proc. Natl. Acad. Sci. USA 1993, 90, 2564–2568. [Google Scholar] [CrossRef] [Green Version]
  50. Pareek, N.; Vivekanand, V.; Agarwal, P.; Saroj, S.; Singh, R.P. Bioconversion to chitosan: A two stage process employing chitin deacetylase from Penicillium oxalicum SAEM-51. Carbohydr. Polym. 2013, 96, 417–425. [Google Scholar] [CrossRef]
  51. Jayakumar, R.; Prabaharan, M.; Nair, S.V.; Tokura, S.; Tamura, H.; Selvamurugan, N. Novel carboxymethyl derivatives of chitin and chitosan materials and their biomedical applications. Prog. Mater Sci. 2010, 55, 675–709. [Google Scholar] [CrossRef]
  52. Bhat, P.; Pawaskar, G.-M.; Raval, R.; Cord-Landwehr, S.; Moerschbacher, B.; Raval, K. Expression of Bacillus licheniformis chitin deacetylase in E. coli pLysS: Sustainable production, purification and characterisation. Int. J. Biol. Macromol. 2019, 131, 1008–1013. [Google Scholar] [CrossRef] [PubMed]
  53. Li, X.; Wang, L.-X.; Wang, X.; Roseman, S. The Chitin Catabolic Cascade in the Marine Bacterium Vibrio Cholerae: Characterization of a Unique Chitin Oligosaccharide Deacetylase. Glycobiology 2007, 17, 1377–1387. [Google Scholar] [CrossRef] [PubMed]
  54. Yang, G.; Hou, X.; Lu, J.; Wang, M.; Wang, Y.; Huang, Y.; Liu, Q.; Liu, S.; Fang, Y. Enzymatic modification of native chitin and chitin oligosaccharides by an alkaline chitin deacetylase from Microbacterium esteraromaticum MCDA02. Int. J. Biol. Macromol. 2022, 203, 671–678. [Google Scholar] [CrossRef]
  55. Tsigos, I.; Bouriotis, V. Purification and Characterization of Chitin Deacetylase from Colletotrichum lindemuthianum (∗). J. Biol. Chem. 1995, 270, 26286–26291. [Google Scholar] [CrossRef] [Green Version]
  56. Chai, J.; Hang, J.; Zhang, C.; Yang, J.; Wang, S.; Liu, S.; Fang, Y. Purification and characterization of chitin deacetylase active on insoluble chitin from Nitratireductor aquimarinus MCDA3-3. Int. J. Biol. Macromol. 2020, 152, 922–929. [Google Scholar] [CrossRef]
  57. Porter, N.J.; Christianson, D.W. Structure, mechanism, and inhibition of the zinc-dependent histone deacetylases. Curr. Opin. Struct. Biol. 2019, 59, 9–18. [Google Scholar] [CrossRef]
  58. Canon, N.; Schein, C.; Chen, X.; Pozzoli, M.; Pathy, V.; Dreskin, S. Alanine scans of IgE-binding to linear epitopes of Ara h 2 reveal critical amino acids. J. Allergy Clin. Immunol. 2021, 147, AB89. [Google Scholar] [CrossRef]
  59. Matthews, B.W.; Nicholson, H.; Becktel, W.J. Enhanced protein thermostability from site-directed mutations that decrease the entropy of unfolding. Proc. Natl. Acad. Sci. USA 1987, 84, 6663–6667. [Google Scholar] [CrossRef] [Green Version]
  60. Fernández-Lucas, J. New Insights on Enzyme Stabilization for Industrial Biocatalysis. ACS Sustain. Chem. Eng. 2021, 9, 15073–15074. [Google Scholar] [CrossRef]
  61. Folch, B.; Dehouck, Y.; Rooman, M. Thermo- and Mesostabilizing Protein Interactions Identified by Temperature-Dependent Statistical Potentials. Biophys. J. 2010, 98, 667–677. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Bradford, M.M. A rapid and sensitive method for the quantitation of microgram quantities of protein utilizing the principle of protein-dye binding. Anal. Biochem. 1976, 72, 248–254. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Protein sequence alignment of SbCDA with PesCDA (Genbank: A0A1L3THR9) [34], PdaA (Genbank: AFQ56715) [25], ClCDA (Genbank: Q6DWK3) [35], AnCDA (Genbank: XP_659456) [30], PgtCDA (Genbank: XP_003323413) [36], and AnCDA2 (Genbank: ACF22101.1). Residues D and H playing catalytic roles are marked with purple triangles; the remaining 19 conserved sites are marked with blue dots.
Figure 1. Protein sequence alignment of SbCDA with PesCDA (Genbank: A0A1L3THR9) [34], PdaA (Genbank: AFQ56715) [25], ClCDA (Genbank: Q6DWK3) [35], AnCDA (Genbank: XP_659456) [30], PgtCDA (Genbank: XP_003323413) [36], and AnCDA2 (Genbank: ACF22101.1). Residues D and H playing catalytic roles are marked with purple triangles; the remaining 19 conserved sites are marked with blue dots.
Molecules 28 00113 g001
Figure 2. Biochemical characteristics of SbCDA. (A) Effects of temperature on the activity of SbCDA. Conditions for stability measurement: Tris-HCl buffer (10 mM, pH 7.0), 6 h. (B) Effects of temperature on the thermostability of SbCDA. (C) Effects of pH on the activity of SbCDA, citrate buffer (10 mM, pH 3.0–6.0), phosphate buffer (10 mM, pH 6.0–7.0), Tris-HCl buffer (10 mM, pH 7.0–9.0), and glycine-NaOH buffer (10 mM, pH 9.0–10.0). Reaction temperature: 40 °C. (D) Effects of pH on the stability in citrate buffer (10 mM, pH 3.0–6.0), phosphate buffer (10 mM, pH 6.0–7.0), Tris-HCl buffer (10 mM, pH 7.0–9.0), and glycine-NaOH buffer (10 mM, pH 9.0–10.0). Incubation conditions: 40 °C, 12 h. Using GlcNAc of 10 g/L as substrate throughout all measurements. All measurements were determined in triplicate, and the error bars indicate the standard deviations (n = 3).
Figure 2. Biochemical characteristics of SbCDA. (A) Effects of temperature on the activity of SbCDA. Conditions for stability measurement: Tris-HCl buffer (10 mM, pH 7.0), 6 h. (B) Effects of temperature on the thermostability of SbCDA. (C) Effects of pH on the activity of SbCDA, citrate buffer (10 mM, pH 3.0–6.0), phosphate buffer (10 mM, pH 6.0–7.0), Tris-HCl buffer (10 mM, pH 7.0–9.0), and glycine-NaOH buffer (10 mM, pH 9.0–10.0). Reaction temperature: 40 °C. (D) Effects of pH on the stability in citrate buffer (10 mM, pH 3.0–6.0), phosphate buffer (10 mM, pH 6.0–7.0), Tris-HCl buffer (10 mM, pH 7.0–9.0), and glycine-NaOH buffer (10 mM, pH 9.0–10.0). Incubation conditions: 40 °C, 12 h. Using GlcNAc of 10 g/L as substrate throughout all measurements. All measurements were determined in triplicate, and the error bars indicate the standard deviations (n = 3).
Molecules 28 00113 g002
Figure 3. Effects of various chemical agents on enzyme activity of SbCDA. CK represents the control group with nothing added. The incubation time of the enzyme with metal ions was 1 h at 37 °C. The reaction time for the enzyme with GlcNAc was 6 h in Tris-HCl buffer (10 mM, pH = 7.0) at 40 °C. All measurements were determined in triplicate, and the error bars indicate the standard deviations (n = 3).
Figure 3. Effects of various chemical agents on enzyme activity of SbCDA. CK represents the control group with nothing added. The incubation time of the enzyme with metal ions was 1 h at 37 °C. The reaction time for the enzyme with GlcNAc was 6 h in Tris-HCl buffer (10 mM, pH = 7.0) at 40 °C. All measurements were determined in triplicate, and the error bars indicate the standard deviations (n = 3).
Molecules 28 00113 g003
Figure 4. Determination of substrate preference of SbCDA. Measured under standard conditions with 10 mg/mL of substrates. 100% activity was used for the best-performing substrate. All measurements were determined in triplicate, and the error bars indicate the standard deviations (n = 3).
Figure 4. Determination of substrate preference of SbCDA. Measured under standard conditions with 10 mg/mL of substrates. 100% activity was used for the best-performing substrate. All measurements were determined in triplicate, and the error bars indicate the standard deviations (n = 3).
Molecules 28 00113 g004
Figure 5. Relative enzyme activity after single-point mutation. Colors indicate the degree of change in enzyme activity (increase or decrease). The activity of SbCDA was defined as 100%. The activity of mutants was obtained by comparison with SbCDA. The redder the color, the more the enzyme activity was changed, and blue represents the opposite. Measurements were conducted under standard conditions using 10 mg/mL of GlcNAc. All measurements were determined in triplicate, and the error bars indicate the standard deviations (n = 3).
Figure 5. Relative enzyme activity after single-point mutation. Colors indicate the degree of change in enzyme activity (increase or decrease). The activity of SbCDA was defined as 100%. The activity of mutants was obtained by comparison with SbCDA. The redder the color, the more the enzyme activity was changed, and blue represents the opposite. Measurements were conducted under standard conditions using 10 mg/mL of GlcNAc. All measurements were determined in triplicate, and the error bars indicate the standard deviations (n = 3).
Molecules 28 00113 g005
Figure 6. SbCDA three-dimensional structure model and molecular docking results.
Figure 6. SbCDA three-dimensional structure model and molecular docking results.
Molecules 28 00113 g006
Figure 7. Molecular structure of SbCDA. (A) Leu62; (B) Thr63; (C) Leu79; (D) Thr87; (E) Phe88; (F) Leu208.
Figure 7. Molecular structure of SbCDA. (A) Leu62; (B) Thr63; (C) Leu79; (D) Thr87; (E) Phe88; (F) Leu208.
Molecules 28 00113 g007
Figure 8. Relative enzyme activity of the mutants after saturation mutation, with the original enzyme activity as 100%. The colors represent the level of enzyme activity. Red means the enzyme activity is higher than the original protein. The redder the color, the higher the enzyme activity, and blue represents the opposite.
Figure 8. Relative enzyme activity of the mutants after saturation mutation, with the original enzyme activity as 100%. The colors represent the level of enzyme activity. Red means the enzyme activity is higher than the original protein. The redder the color, the higher the enzyme activity, and blue represents the opposite.
Molecules 28 00113 g008
Figure 9. Biochemical characteristics of SbCDA63G. (A) Effects of temperature on the activity of SbCDA63G. Conditions for stability measurement: Tris-HCl buffer (10 mM, pH 7.0), incubation for 6 h. (B) Effects of temperature on the thermostability of SbCDA63G. (C) Effects of pH on the activity of SbCDA63G: citrate buffer (10 mM, pH 3.0–6.0), phosphate buffer (10 mM, pH 6.0–7.0), Tris-HCl buffer (10 mM, pH 7.0–9.0), and glycine-NaOH buffer (10 mM, pH 9.0–10.0). Reaction temperature: 40 °C. (D) Effects of pH on the stability in citrate buffer (10 mM, pH 3.0–6.0), phosphate buffer (10 mM, pH 6.0–7.0), Tris-HCl buffer (10 mM, pH 7.0–9.0), and glycine-NaOH buffer (10 mM, pH 9.0–10.0). Incubation conditions: 40 °C, incubation for 12 h. Using GlcNAc of 10 g/L as substrate throughout all measurements. All measurements were determined in triplicate, and the error bars indicate the standard deviations (n = 3).
Figure 9. Biochemical characteristics of SbCDA63G. (A) Effects of temperature on the activity of SbCDA63G. Conditions for stability measurement: Tris-HCl buffer (10 mM, pH 7.0), incubation for 6 h. (B) Effects of temperature on the thermostability of SbCDA63G. (C) Effects of pH on the activity of SbCDA63G: citrate buffer (10 mM, pH 3.0–6.0), phosphate buffer (10 mM, pH 6.0–7.0), Tris-HCl buffer (10 mM, pH 7.0–9.0), and glycine-NaOH buffer (10 mM, pH 9.0–10.0). Reaction temperature: 40 °C. (D) Effects of pH on the stability in citrate buffer (10 mM, pH 3.0–6.0), phosphate buffer (10 mM, pH 6.0–7.0), Tris-HCl buffer (10 mM, pH 7.0–9.0), and glycine-NaOH buffer (10 mM, pH 9.0–10.0). Incubation conditions: 40 °C, incubation for 12 h. Using GlcNAc of 10 g/L as substrate throughout all measurements. All measurements were determined in triplicate, and the error bars indicate the standard deviations (n = 3).
Molecules 28 00113 g009
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Yin, L.; Wang, Q.; Sun, J.; Mao, X. Expression and Molecular Modification of Chitin Deacetylase from Streptomyces bacillaris. Molecules 2023, 28, 113. https://doi.org/10.3390/molecules28010113

AMA Style

Yin L, Wang Q, Sun J, Mao X. Expression and Molecular Modification of Chitin Deacetylase from Streptomyces bacillaris. Molecules. 2023; 28(1):113. https://doi.org/10.3390/molecules28010113

Chicago/Turabian Style

Yin, Lili, Qi Wang, Jianan Sun, and Xiangzhao Mao. 2023. "Expression and Molecular Modification of Chitin Deacetylase from Streptomyces bacillaris" Molecules 28, no. 1: 113. https://doi.org/10.3390/molecules28010113

Article Metrics

Back to TopTop