Skip to main content
  • Research article
  • Open access
  • Published:

Consensus design for improved thermostability of lipoxygenase from Anabaena sp. PCC 7120

Abstract

Background

Lipoxygenase (LOX) from Anabaena sp. PCC 7120 (Ana-rLOX) offers important applications in the food industry, especially for improving aroma and dough rheological properties. However, industrial applications of LOXs have been limited by their poor thermostability. Herein, we report a bioinformatics-based consensus concept approach for the engineering of thermostable Ana-rLOX.

Results

A series of mutations (N130D, G260A, S437T, N130D/G260Q, N130D/S437Y) showed higher thermostability and activity than the wild-type enzyme. Thus, N130D/G260Q exhibited a 6.6-fold increase in half-life and 2.45 °C increase in unfolding temperature; N130D/S437Y showed a 10 °C increase in optimal temperature. The secondary structure did not change much that contributed to improved thermostability were investigated in detail using circular dichroism. Homology modeling suggested that enhanced thermostability and specific activity may result from favorable hydrophobic interactions.

Conclusions

A series of mutations were achieved, showing higher thermostability and activity than the wild-type enzyme by semi-rational mutagenesis with limited structure information. Our findings provide important new insights into molecular modifications aimed at improving Ana-rLOX thermostability and activity.

Background

Lipoxygenases (LOXs; EC1.13.11.12) are a family of non-heme iron-containing dioxygenases that catalyze regio- and stereospecific dioxygenation of polyunsaturated fatty acids(PUFAs) containing a 1-cis,4-cis-pentadiene structure to form fatty acid hydroperoxides [1]. LOX-catalyzed oxidative coupling reactions are of interest to the food industry. They are applied in bread-making to bleach carotenoids and to improve dough rheology [2, 3]. In addition, LOX also played an important role in aroma compound formation. Because of their unique properties, LOXs have been widely utilized in the food, pharmaceutical [4, 5] and textile industries [6, 7]. However, industrial applications of LOXs have been relatively limited because of their poor thermostability, it is essential to generate LOXs with enhanced activity and thermostability.

Numerous protein engineering strategies have been attempted to improve protein stability through modification of various structural features [8]. Directed evolution, rational design, and semi-rational design methods have been used successfully to overcome certain limitations of enzymes and to improve thermostability [9,10,11]. Although some of these approaches have led to improving LOX thermostability [12], protein engineering aimed at strengthening thermostability and catalytic activity remains a challenging and exciting field [13]. This is particularly true when structural and mechanistic information about an enzyme is very limited, making it difficult to rationally predict the function of targeted amino acid residues [14]. “Consensus concept” is a rapidly developing theoretical method in protein engineering [15], which requires only protein family sequence information. Consensus concept assumes that the amino acid appearing most frequently among homologues at a specific position, contributes more than other residues at that position to enzyme stability. The method has been applied successfully to ameliorate protein thermostability [16,17,18].

In our previous studies, we cloned, expressed, and characterized the Ana-rLOX gene from Anabaena sp. PCC 7120, and this enzyme had also been successfully expressed in Bacillus subtilis(generally recommended as safe, GRAS) [19], so that Ana-rLOX could be used in food processing safely. The cloned enzyme (Ana-rLOX) had high specific activity but low stability, and was therefore unsuitable for many industrial applications [12]. Directed evolution is not a viable approach because there is no efficient high-throughput screening method for LOX [20]. In the present study, we employed sequence information about the LOX family to determine mutation sites on the basis of the consensus concept and constructed “small but smart” saturated mutant libraries for selection of Ana-rLOX mutant enzymes with enhanced thermostability. This approach allowed us to generate a series of mutants with greatly improved thermostability.

Methods

Materials

Ni2+-nitrilotriacetate (Ni-NTA) was purchased from Qiagen (Hilden, Germany). Linoleic acid was purchased from Sigma (Steinheim, Germany). PrimeSTAR DNA polymerase, enzyme Dpn I, and the genomic DNA purification kit were from TaKaRa (Dalian, China). Isopropyl β-D-l-thiogalactopyranoside (IPTG) and ampicillin were from Sangon Biotech (Shanghai, China). All compounds were of reagent grade or higher.

Cloning and expression strains were purchased from TaKaRa. E. coli BL21(DE3) and expression vector pET-32a were used for protein expression.

Analysis of consensus amino acids

To identify key sites for mutagenesis, a simple approach is to align the target enzyme with its closely related homologs. Amino acid sequences common to the LOX family were downloaded from the pfam [21] and GenBank databases. Similarities between Ana-rLOX and LOX family proteins were analyzed using the Basic Local Alignment Search Tool (BLAST; NCBI). In contrast to other sequence-selection methods and because Ana-rLOX has low sequence identity with other LOX family proteins, we selected the 100-most similar LOX sequences. The selected sequences were aligned using ClustalX2, generating a phylogenetic tree by Neighbour Joining approach, which was then used to group the sequences into subsets. The most relevant sequences were chosen for analysis and were set to represent the main branches of the phylogenetic tree. The success of this approach depends on the phylogenetic diversity of the available sequence set [22]. The consensus sequence was calculated by ESpript [23] and either consensus positions or most abundant positions were determined. On the basis of the consensus result, non-consensus amino acids were substituted by consensus amino acids using site-directed saturation mutagenesis.

Construction of saturation mutagenesis libraries

The plasmid pET-32a/Ana-rLOX encoding the LOX gene from Anabaena sp. PCC 7120 was used to construct the libraries [12], Ana-rLOX gene is 1365 bp. Mutants were prepared by the Quik Change® site-directed mutagenesis method [24] with primers containing NNK/MNN (K = G/T,M = A/C,N = A/C/G/T) degeneracy at target sites. Primers are listed in Additional file 1: Table S1.

Amplification was performed using the following temperature settings: 2 min at 98 °C; 30 cycles of 10 s at 98 °C, 15 s at 55 °C, and 7 min at 72 °C; final 10-min extension at 72 °C. PCR products were digested with Dpn I to remove the parent plasmid.

The created libraries were transformed into E. coli BL21(DE3) and selected for ampicillin resistance. Clones were picked in 96-well plates containing 200 μL Luria-Bertani (LB) medium with 100 μg/mL ampicillin. After growth at 37 °C overnight, the cultures were used to inoculate fresh medium in a new 96-well plate and incubated at 37 °C. IPTG was added to induce expression of the target protein at 16 °C. Due to the use of degenerate primers (NNK), each saturation mutagenesis library must have contained at least 200 colonies to overlay 20 kinds of amino acid codons. Screening was performed as described previously [12]. Selected positive clones were tested independently four more times. Confirmed positive clones were sequenced, and enzymes were expressed, purified, and characterized.

Enzyme production and purification

For a detailed characterization of Ana-rLOX and its variants, E. coli BL21(DE3) transformed with expression plasmids was used. For enzyme production, 2% of overnight culture was inoculated in 100 mL LB medium containing 100 mg/mL ampicillin, and incubated at 37 °C with continuous shaking (180 rpm) until late logarithmic phase. Target protein expression was induced by addition of IPTG (final concentration 100 mg/mL). Protein induction was performed at 16 °C for 16 h. Cells were harvested by centrifugation (8000×g, 10 min, 4 °C), resuspended in 10 mL phosphate-buffered saline (PBS; 300 mM NaCl, 50 mM phosphate buffer, pH 8.0), and the obtained cell pellets were subjected to sonication. The crude extract was centrifuged as above to remove cell debris.

Crude enzyme was loaded into Ni-NTA resin to purify the protein, utilizing the His-tag encoded by pET-32a. SDS-PAGE analysis was performed on a 12% running gel [25], and resolved proteins were visualized by Coomassie Brilliant Blue G-250 staining. Protein concentration was measured by the Bradford method [26] with bovine serum albumin as the standard.

Enzyme assays

Enzyme activity assays were performed as described previously [27]. Ana-rLOX activity was assayed at 30 °C using linoleic acid as substrate, and Tris-HCl (pH 9.0) as buffer. One unit of activity was defined as the quantity of enzyme required to synthesize 1 μM hydroperoxide, with molar extinction coefficient (ε) = 25,000 L/(mol·cm) [28].

Hydroperoxidation was determined at 234 nm in a spectrophotometer (UV-2450; Shimadzu Co., Kyoto, Japan). Enzyme activity was calculated using the initial linear part of the curve.

Enzyme stability

Kinetic stability was assessed by incubating the purified enzyme (2.0 mg/mL) in a 50 °C water bath with sampling at regular intervals. Residual activity was measured and compared to that of the untreated enzyme. Thermostability was evaluated at 50 °C. Half-life (t1/2) was calculated after fitting experimental data to the equation ln Y = −kdX + b, assuming first order inactivation (t1/2 = ln 2/ kd). To construct the melting curves and determine the melting temperature (Tm), samples were heated at 1 °C per min from 20 to 90 °C and the circular dichroism (CD) ellipticity signal at 222 nm, which showed the maximal change with temperature, was monitored.

Circular dichroism analysis

CD analysis was performed using a spectropolarimeter (model J-810, Jasco, Tokyo, Japan). Enzyme samples were dissolved in 10 mM PBS, pH 7.4, at a concentration of 0.1–0.15 mg/mL and loaded on a quartz cuvette (path length 1 cm). The spectrum of a buffer blank was subtracted. Helical and sheet contents of proteins were estimated using the online program DICHROWEB [29].

Homology modeling

To analyze the mutated residues, the structures of Ana-rLOX and its mutants were simulated by Phyre2 software [30]. Lipoxigenase from Cyanothece sp. PCC 8801 (PDB:5EK8) [31] was used as template.

Results

Mutation design guided by the consensus concept

The consensus concept is a bioinformatics-based protein design method that replaces specific amino acids with the amino acid that most frequently occurs at that position in the protein family [32, 33]. The amino acid sequence of Ana-rLOX was used to identify homologous sequences in the GenBank database. Sequence identity is critical to successfully determine the consensus sequence. Sequences with a close phylogenetic relationship with Ana-rLOX were selected, but those with > 95% identity were removed to avoid bias caused by repetition of similar sequences [8]. Therefore, sequences with identities between 30 to 90% were chosen for multiple sequence alignment (Fig. 1), and most of the stable amino acids in evolution were selected.

Fig. 1
figure 1

Phylogenetic tree of LOX enzymes. These sequences were chosen for multiple sequence alignment

To reduce number of the candidated amino acids and improve quality of the mutagenesis libraries, homologous model of Ana-rLOX was utilized in selecting mutation sites. First, residues within 5 Å of the active site were excluded to avoid compromising enzyme activity. Second, mutations that would introduce hydrophobic amino acids on the protein surface or hydrophilic amino acids in the protein core were discarded. Third, mutations that would destroy existing hydrogen-bonds, salt-bridges, or other stabilizing interactions were also avoided. Consequently, ten positions was left for constructing saturation mutagenesis libraries with which to further investigate changes in thermal stability (Fig. 2).

Fig. 2
figure 2

Protein sequence of Ana-rLOX. Ten positions indicating in bold were left for constructing saturated mutagenesis libraries,and metal liganding amino acids were labeled by triangle()

Characterization of mutants with improved thermostability

The ten selected residues were subjected to saturation mutagenesis and screening revealed several mutants with improved thermostability. Combinational mutagenesis experiments were performed in the next round by selecting the best mutant, N130D, as template; N130D/S437Y and N130D/G260Q were identified as providing further stability enhancement.

All the above mutants presented higher optimum temperatures (Table 1). For example, mutants S437T and N130D/S437Y showed highest activity at 45 °C, which is 10 °C higher than the wild-type enzyme, and thus may be suitable for industrial applications at such temperatures. At the same time, the half-lives (t1/250°C) defined by the temperature at which the enzyme loses 50% of activity, increased from 2.3 min for wild-type to 9.3 min and 11.0 min for S437T and N130D/S437Y, respectively. The most thermostable combinational mutant, N130D/G260Q, exhibited a 6.6-fold longer half-life than the wild-type. Tm values determined by CD spectroscopy (Additional file 1: Figure S2), revealed a modest increase of 1.4 °C in mutant N130D compared to wild-type. To further increase thermostability, we tried to combine this mutation with S437Y and G260Q and obtained increases of 2.1–2.5 °C. These results indicate that combinations of beneficial amino acid substitutions correlated with significant structural changes, which contributed to improved thermostability.

Table 1 Thermal stability of mutants

In addition to improved thermostability, all mutant enzymes exhibited ameliorations in specific activities (Table 2). The specific activity of mutants S437T and N130D/G260Q was 1.7 and 2-fold higher than wild-type, respectively. This finding indicates that both thermostability and specific activity can be improved at the same time by the consensus method.

Table 2 Characterization of mutations

To compare the catalytic reactions between wild-type and mutants, kinetic parameters were determined (Table 2). S437T displayed no changes in Km and an increase of 288% in kcat compared to wild-type, indicating that the mutant did not change its binding affinity for the substrate and had higher specific activity. Interestingly, the two double mutants N130D/S437Y and N130D/G260Q exhibited a 73% and 77% decrease in Km, respectively, indicating higher relative affinity for the substrate, and a consequent 290–400% increase in catalytic efficiency (kcat/ Km). This result indicates that enhanced thermal tolerance was likely accompanied by improved catalytic activity.

Structural analysis of the mutations

The CD spectra of the thermostable variants were analyzed to determine whether the mutations affected the protein’s secondary structure (Additional file 1: Figure S2). Results showed that when the mutation was introduced at the N130 position, the secondary structure changed almostly nothing. When mutation S437T or G260Q was introduced, the β-sheet content increased to 12.4% and 16.1%, respectively (Table 3), revealing that the combing mutations influenced the secondary structure of the enzyme more than one point mutations.

Table 3 Secondary structure prediction by circular dichroism

To gain deeper insight into the structural basis of thermal stability, a molecular model of the mutant was constructed. Substitution of Gly with hydrophobic residue Ala at position 260 resulted in enhanced hydrophobic interactions. The structure of G260A showed that position 260 was located in the surface loop region of the protein (Fig. 3). Stabilization of loops is known to play an important role in protein stability [34].

Fig. 3
figure 3

G260A substitution was located on the loop region of Ana-rLOX. Substitution of Gly with hydrophobic residue Ala at position 260 resulted in enhanced hydrophobic interactions

The substitution G260A has a similar effect as S437T, introducing one extra methyl group into Ana-rLOX that enhanced the hydrophobic interactions within the protein. Interestingly, the combinational mutants exhibited greater improvement when the mutation N130D was accompanied by substitutions G260Q and S437Y but not G260A or S437T.

Discussion

Consensus concept is an appealing strategy for increasing the stability of proteins, as it relies only on sequence information [18] and, particularly, multiple sequences alignment. Nevertheless, even this method can only help determine the candidate residues to be mutated, but cannot predict the specific amino acid that could improve the enzyme. In this study, we investigated the difference between site-directed mutagenesis and saturation mutagenesis guided by the consensus concept. Half-lives of mutations N130D, S437T, and G260A were all significantly higher than in the wild-type (Table 4). We show that mutations introduced at G260 and S437 were not in accord with the consensus result, only mutation N130D was the “consensus amino acid”. We proposed that consensus approach could predict the position deserved to be mutated [16], and saturation mutagenesis was very efficient at improving the success rate of the experiment..

Table 4 Comparative effects of site-directed mutagenesis and saturation mutagenesis

Three mutants (G260A, S437T, and N130D) produced on the basis of a structure-guided consensus method displaying varying degrees of improvements in thermostability (Table 2). The structure of G260A based on homology modeling showed that position 260 was located in the surface loop region of the protein (Fig. 3). Stabilization of loops, which plays an important role in protein stability, can be enhanced by engineering loop residues [35]. Substitution of Gly with Ala at position 260 resulted in enhanced hydrophobic interactions, promoting the formation of α-helix regions (Fig. 3), and improving stability of Ana-LOX through a more stable secondary structure. A effect was observed for S437T, whose half-life was also 4-fold higher than wild-type. Here, introduction of an extra methyl group into Ana-LOX which contributed increasing hydrophobic interaction (Fig. 4). These findings indicated us that enhanced enzyme thermostability depends mainly on favorable hydrophobic interactions.

Fig. 4
figure 4

S437T mutation introduced one extra methyl group into Ana-rLOX, that enhanced the hydrophobic interactions within the protein

Interestingly, all of our consensus mutations improved both thermostability and specific activity of Ana-LOX relative to wild-type. Although the consensus method has been found to be effective for increasing enzyme stability, such mutations are successful in only about 50% of cases [18]. As a matter of fact, thermostability and specific activity are partially independent properties, even though they can be optimized together in a given enzyme [36]. In a previous study using site-directed mutagenesis based on computer-aided rational design, the mutant V421A/V40A showed a 1.13-fold increase in thermostability and 80.07% increase in specific activity [37]. In the present study using consensus concept with saturation mutagenesis, the mutant N130D/G260Q showed a 6.6-fold increase in half-life at 50 °C and 2.0-fold increase in specific activity relative to Ana-rLOX. Thus, the consensus method proved efficient at screening Ana-rLOX mutants for enhanced thermostability and specific activity. Moreover, it allowed for significant increases in thermostability in comparison with alternative techniques such as rational design.

Conclusion

Ana-rLOX mutants showing significant improvements in both thermostability and specific activity were produced by semi-rational mutagenesis based on sequence alignment and structural analysis. Mutant N130D/G260Q exhibited a 6.6-fold increase in half-life, and mutant N130D/S437Y showed a 10 °C increase in optimal temperature. Additionally, the catalytic efficiency (kcat/Km) was also improved. Our findings provide important new insights into molecular modifications aimed at improving Ana-LOX thermostability and activity.

References

  1. Kuhn H, Banthiya S, van Leyen K. Mammalian lipoxygenases and their biological relevance. Biochim Biophys Acta. 2015;1851(4):308–30.

    Article  CAS  Google Scholar 

  2. Fukushige H, Hildebrand DF. A simple and efficient system for green note compound biogenesis by use of certain lipoxygenase and hydroperoxide lyase sources. J Agric Food Chem. 2005;53(17):6877–82.

    Article  CAS  Google Scholar 

  3. Zhang C, Zhang S, Lu Z, Bie X, Zhao H, Wang X, Lu F. Effects of recombinant lipoxygenase on wheat flour, dough and bread properties. Food Res Int. 2013;54(1):26–32.

    Article  Google Scholar 

  4. Lu W, Zhao X, Xu Z, Dong N, Zou S, Shen X, Huang J. Development of a new colorimetric assay for lipoxygenase activity. Anal Biochem. 2013;441(2):162–8.

    Article  CAS  Google Scholar 

  5. Wu Z, Robinson DS, Hughes RK, Casey R, Hardy D, West SI. Co-oxidation of beta-carotene catalyzed by soybean and recombinant pea lipoxygenases. J Agric Food Chem. 1999;47(12):4899–906.

    Article  CAS  Google Scholar 

  6. Xu Z, Liu S, Lu X, Rao S, Kang Z, Li J, Wang M, Chen J. Thermal inactivation of a recombinant lipoxygenase from Pseudomonas aeruginosa BBE in the absence and presence of additives. J Sci Food Agric. 2014;94(9):1753–7.

    Article  CAS  Google Scholar 

  7. Wang XM, Lu FX, Zhang C, Lu YJ, Bie XM, Xie YJ, Lu ZX. Effects of Recombinated Anabaena sp. lipoxygenase on the protein component and dough property of wheat flour. J Agric Food Chem. 2014;62(40):9885–92.

    Article  CAS  Google Scholar 

  8. Vazquez-Figueroa E, Chaparro-Riggers J, Bommarius AS. Development of a thermostable glucose dehydrogenase by a structure-guided consensus concept. Chembiochem. 2007;8(18):2295–301.

    Article  CAS  Google Scholar 

  9. Feng X, Tang H, Han B, Lv B, Li C. Enhancing the Thermostability of β-Glucuronidase by rationally redesigning the catalytic domain based on sequence alignment strategy. Ind Eng Chem Res. 2016;55(19):5474–83.

    Article  CAS  Google Scholar 

  10. Tian J, Wang P, Huang L, Chu X, Wu N, Fan Y. Improving the thermostability of methyl parathion hydrolase from Ochrobactrum sp. M231 using a computationally aided method. Appl Microbiol Biotechnol. 2013;97(7):2997–3006.

    Article  CAS  Google Scholar 

  11. Johannes TW, Zhao HM. Directed evolution of enzymes and biosynthetic pathways. Curr Opin Microbiol. 2006;9(3):261–7.

    Article  CAS  Google Scholar 

  12. Guo F, Zhang C, Bie X, Zhao H, Diao H, Lu F, Lu Z. Improving the thermostability and activity of lipoxygenase from Anabaena sp. PCC 7120 by directed evolution and site-directed mutagenesis. J Mol Catal B Enzym. 2014;107:23–30.

    Article  CAS  Google Scholar 

  13. Yu H, Huang H. Engineering proteins for thermostability through rigidifying flexible sites. Biotechnol Adv. 2014;32(2):308–15.

    Article  CAS  Google Scholar 

  14. Liang C, Gui X, Zhou C, Xue Y, Ma Y, Tang SY. Improving the thermoactivity and thermostability of pectate lyase from Bacillus pumilus for ramie degumming. Appl Microbiol Biotechnol. 2015;99(6):2673–82.

    Article  CAS  Google Scholar 

  15. Polizzi KM, Chaparro-Riggers JF, Vazquez-Figueroa E, Bommarius AS. Structure-guided consensus approach to create a more thermostable penicillin G acylase. Biotechnol J. 2006;1(5):531–6.

    Article  CAS  Google Scholar 

  16. Aerts D, Verhaeghe T, Joosten H-J, Vriend G, Soetaert W, Desmet T. Consensus engineering of sucrose phosphorylase: the outcome reflects the sequence input. Biotechnol Bioeng. 2013;110(10):2563–72.

    Article  CAS  Google Scholar 

  17. Bvd B, Eijsink VG. Selection for increased protein stability. Curr Opin Biotechnol. 2002;13(4):333–7.

    Article  Google Scholar 

  18. Sullivan BJ, Nguyen T, Durani V, Mathur D, Rojas S, Thomas M, Syu T, Magliery TJ. Stabilizing proteins from sequence statistics: the interplay of conservation and correlation in triosephosphate isomerase stability. J Mol Biol. 2012;420(4–5):384–99.

    Article  CAS  Google Scholar 

  19. Zhang C, Tao T, Ying Q, Zhang D, Lu F, Bie X, Lu Z. Extracellular production of lipoxygenase from Anabaena sp. PCC 7120 in Bacillus subtilis and its effect on wheat protein. Appl Microbiol Biotechnol. 2012;94(4):949–58.

    Article  CAS  Google Scholar 

  20. Li SF, Xu JY, Bao YJ, Zheng HC, Song H. Structure and sequence analysis-based engineering of pullulanase from Anoxybacillus sp. LM18-11 for improved thermostability. J Biotechnol. 2015;210:8–14.

    Article  CAS  Google Scholar 

  21. Finn RD, Coggill P, Eberhardt RY, Eddy SR, Mistry J, Mitchell AL, Potter SC, Punta M, Qureshi M, Sangrador-Vegas A, et al. The Pfam protein families database: towards a more sustainable future. Nucleic Acids Res. 2016;44(D1):D279–85.

    Article  CAS  Google Scholar 

  22. Jackel C, Bloom JD, Kast P, Arnold FH, Hilvert D. Consensus protein design without phylogenetic bias. J Mol Biol. 2010;399(4):541–6.

    Article  Google Scholar 

  23. Robert X, Gouet P. Deciphering key features in protein structures with the new ENDscript server. Nucleic Acids Res. 2014;42(Web Server issue):W320–4.

    Article  CAS  Google Scholar 

  24. Zheng L: An efficient one-step site-directed and site-saturation mutagenesis protocol. Nucleic Acids Res 2004, 32(14):e115-e115.

    Article  Google Scholar 

  25. Laemmli UK. Cleavage of structural proteins during the assembly of the head of bacteriophage T4. Nature. 1970;227(5259):680–5.

    Article  CAS  Google Scholar 

  26. Bradford MM. A rapid and sensitive method for the quantitation of microgram quantities of protein utilizing the principle of protein-dye binding. Anal Biochem. 1976;72:248–54.

    Article  CAS  Google Scholar 

  27. Lu X, Liu S, Zhang D, Zhou X, Wang M, Liu Y, Wu J, Du G, Chen J. Enhanced thermal stability and specific activity of Pseudomonas aeruginosa lipoxygenase by fusing with self-assembling amphipathic peptides. Appl Microbiol Biotechnol. 2013;97(21):9419–27.

    Article  CAS  Google Scholar 

  28. Hughes RK, Wu ZC, Robinson DS, Hardy D, West SI, Fairhurst SA, Casey R. Characterization of authentic recombinant pea-seed lipoxygenases with distinct properties and reaction mechanisms. Biochem J. 1998;333:33–43.

    Article  CAS  Google Scholar 

  29. Whitmore L, Wallace BA. DICHROWEB, an online server for protein secondary structure analyses from circular dichroism spectroscopic data. Nucleic Acids Res. 2004;32(Web Server):W668–73.

    Article  CAS  Google Scholar 

  30. Kelley LA, Mezulis S, Yates CM, Wass MN, Sternberg MJE. The Phyre2 web portal for protein modeling, prediction and analysis. Nat Protoc. 2015;10(6):845–58.

    Article  CAS  Google Scholar 

  31. Newie J, Andreou A, Neumann P, Einsle O, Feussner I, Ficner R. Crystal structure of a lipoxygenase from Cyanothece sp may reveal novel features for substrate acquisition. J Lipid Res. 2016;57(2):276–87.

    Article  CAS  Google Scholar 

  32. Anbar M, Gul O, Lamed R, Sezerman UO, Bayer EA. Improved thermostability of Clostridium thermocellum endoglucanase Cel8A by using consensus-guided mutagenesis. Appl Environ Microbiol. 2012;78(9):3458–64.

    Article  CAS  Google Scholar 

  33. Silva IR, Larsen DM, Jers C, Derkx P, Meyer AS, Mikkelsen JD. Enhancing RGI lyase thermostability by targeted single point mutations. Appl Microbiol Biotechnol. 2013;97(22):9727–35.

    Article  CAS  Google Scholar 

  34. Martinez R, Jakob F, Tu R, Siegert P, Maurer K-H, Schwaneberg U. Increasing activity and thermal resistance of Bacillus gibsonii alkaline protease (BgAP) by directed evolution. Biotechnol Bioeng. 2013;110(3):711–20.

    Article  CAS  Google Scholar 

  35. Martinez CA, Giulietti AM, Rodriguez Talou J. Research advances in plant-made flavivirus antigens. Biotechnol Adv. 2012;30(6):1493–505.

    Article  CAS  Google Scholar 

  36. Lehmann M, Pasamontes L, Lassen SF, Wyss M. The consensus concept for thermostability engineering of proteins. Biochimica Et Biophysica Acta-Protein Structure and Molecular Enzymology. 2000;1543(2):408–15.

    Article  CAS  Google Scholar 

  37. Diao H, Zhang C, Wang S, Lu F, Lu Z. Enhanced Thermostability of lipoxygenase from Anabaena sp. PCC 7120 by site-directed mutagenesis based on computer-aided rational design. Appl Biochem Biotechnol. 2016;178(7):1339–50.

    Article  CAS  Google Scholar 

Download references

Acknowledgements

The authors are grateful to Dr. S. Anderson for English editing of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China [grant numbers 31470095, 31671800].

Availability of data and materials

The datasets used and/or analyzed during the current study are available from the corresponding author on reasonable request.

Author information

Authors and Affiliations

Authors

Contributions

HQ and BX performed experiments and analyzed data; CZ, ZL, XB, HZ contributed materials; FL and G-YY analyzed data and designed experiments; HQ and FL wrote the paper. All authors read and approved the final manuscript.

Corresponding authors

Correspondence to Fengxia Lu or Guang-Yu Yang.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Additional file

Additional file 1:

Table S1. Saturation mutagenesis primers. Figure S1. Purification of wild-type and mutants. M: Marker, 1:wild, 2: N130D, 3: G260A, 4: S437T, 5: N130D/S437Y, 6: N130D/G260Q. Crude enzyme was loaded onto Ni-NTA resin to purify the protein, utilizing the His-tag encoded by pET-32a. SDS-PAGE analysis was performed on a 12% running gel and were visualized by Coomassie Brilliant Blue G-250 staining. Figure S2. Tm values of enzymes determined by differential scanning calorimetry. Figure S3. Circular dichroism of wild-type and mutants. (DOC 320 kb)

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Qian, H., Zhang, C., Lu, Z. et al. Consensus design for improved thermostability of lipoxygenase from Anabaena sp. PCC 7120. BMC Biotechnol 18, 57 (2018). https://doi.org/10.1186/s12896-018-0468-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12896-018-0468-4

Keywords