DOI QR코드

DOI QR Code

Production of Biopharmaceuticals in E. coli: Current Scenario and Future Perspectives

  • Baeshen, Mohammed N. (Department of Biological Sciences, Faculty of Science, King Abdulaziz University) ;
  • Al-Hejin, Ahmed M. (Department of Biological Sciences, Faculty of Science, King Abdulaziz University) ;
  • Bora, Roop S. (Department of Biological Sciences, Faculty of Science, King Abdulaziz University) ;
  • Ahmed, Mohamed M. M. (Department of Biological Sciences, Faculty of Science, King Abdulaziz University) ;
  • Ramadan, Hassan A. I. (Department of Biological Sciences, Faculty of Science, King Abdulaziz University) ;
  • Saini, Kulvinder S. (Department of Biological Sciences, Faculty of Science, King Abdulaziz University) ;
  • Baeshen, Nabih A. (Department of Biological Sciences, Faculty of Science, King Abdulaziz University) ;
  • Redwan, Elrashdy M. (Department of Biological Sciences, Faculty of Science, King Abdulaziz University)
  • Received : 2014.12.30
  • Accepted : 2015.02.13
  • Published : 2015.07.28

Abstract

Escherichia coli is the most preferred microorganism to express heterologous proteins for therapeutic use, as around 30% of the approved therapeutic proteins are currently being produced using it as a host. Owing to its rapid growth, high yield of the product, costeffectiveness, and easy scale-up process, E. coli is an expression host of choice in the biotechnology industry for large-scale production of proteins, particularly non-glycosylated proteins, for therapeutic use. The availability of various E. coli expression vectors and strains, relatively easy protein folding mechanisms, and bioprocess technologies, makes it very attractive for industrial applications. However, the codon usage in E. coli and the absence of post-translational modifications, such as glycosylation, phosphorylation, and proteolytic processing, limit its use for the production of slightly complex recombinant biopharmaceuticals. Several new technological advancements in the E. coli expression system to meet the biotechnology industry requirements have been made, such as novel engineered strains, genetically modifying E. coli to possess capability to glycosylate heterologous proteins and express complex proteins, including full-length glycosylated antibodies. This review summarizes the recent advancements that may further expand the use of the E. coli expression system to produce more complex and also glycosylated proteins for therapeutic use in the future.

Keywords

Introduction

The drug discovery and development process entails the expression of large numbers of recombinant proteins in copious amounts and properly folded as 3D structures. These therapeutic recombinant proteins are generally used for the treatment of various diseases, or are used as a target protein for screening of new and novel drugs. E. coli is one of the most desirable host for the expression of several recombinant proteins owing to its rapid growth rate, easier genetic manipulations, and high level of recombinant protein synthesis rates [63, 65]. In a few instances, very high levels of expression are achieved; that is, up to 30% of total cellular protein. Escherichia coli was the first expression host that was used for manufacturing a biopharmaceutical, which resulted in the regulatory approval of human insulin in 1982 for the treatment of diabetes. The approval of bovine growth hormone (bGH) in 1994 set a new standard for manufacturing heterologous proteins from E. coli for therapeutic use (Table 1). It is even more impressive that both insulin and bovine growth hormone require oxidative protein folding, and, in addition, insulin is a heterodimer. Success achieved with recombinant insulin and growth hormone highlights the versatility and cost-effectiveness of E. coli-based production.

Table 1.All data obtained from corporate websites and http://www.fda.gov/. rh, recombinant human; G-CSF, Granulocyte-colony stimulating factor; EU, European union; US, United States of America.

However, in E. coli, there is a probability of translational errors due to the presence of a large number of rare codons in the heterologous gene(s). In the case of therapeutic proteins, these errors, even at low levels, can cause adverse immunogenic responses in humans. These cellular errors during protein translation may impact the tertiary structure and thus affect the biological activity of the recombinant protein. In this review, we will be focusing on the “desired modulation” of key biochemical parameters required to obtain properly folded and overexpressed biologically active therapeutic proteins in E. coli.

 

Codon Usage in E. coli

In E. coli, codon usage is manifested by the level of cognate aminoacylated tRNAs in the cytosol. Major codons are generally present in genes that are expressed at a high level; on the other hand, rare codons are commonly encountered in genes that are expressed at low levels. Codons that are rare in E. coli are found to be abundant in eukaryotic genes [31]. Expression of heterologous genes harboring rare codons can result in translational errors, due to ribosomal stalling at positions where amino acids coupled to rare codon tRNAs have to be incorporated [43]. In addition, translational errors due to the presence of rare codons in heterologous genes might include amino acid substitutions, frame-shift mutations, or premature termination of translation [31, 69]. Kane et al. [30] had reported inframe, two amino acids “hops” at a rare AGA codon. It had also been reported that protein quality is affected to a great extent by codon bias, due to the incorporation of lysine for arginine at the AGA codon [4, 67]. Hence, expression of recombinant proteins, even at very high levels, is of no use if the quality of the protein is compromised as a result of translational errors. The most problematic rare codons in E. coli are AGA, AGG, CGG, CGA, CGG (arginine), AUA (isoleucine), GGA (glycine), CUA (leucine), CCC (proline), and AAG (lysine) [85]. It has been documented that rare arginine codons, AGG and AGA, occur at frequencies of 0.14% and 0.21%, respectively, in E. coli [31].

Several strategies are employed to circumvent the issue of codon bias in E. coli. One approach is to synthesize the whole gene based on codon usage, which is currently a preferred method to improve the expression of heterologous proteins in E. coli [59, 61]. However, a major drawback is the high cost associated with total gene synthesis. Another approach involves the site-directed mutagenesis of the heterologous gene sequence to generate codons, which reflects the tRNA pool of E. coli. However, this process may become very tedious and expensive, whenever several nucleotides need to be modified. Another strategy requires the co-transformation of E. coli strains with plasmid harboring a gene encoding the tRNA cognate to the rare codons [12]. By enhancing the copy number of the limiting tRNAs, E. coli can be manipulated to match the codon usage frequency in heterologous genes. Several commercial plasmids, such as pRARE, are now available for rare tRNA coexpression in E. coli. Moreover, these plasmids contain the p15A replication origin, which enables their maintenance in the presence of the ColE1 replication origin in the E. coli expression vectors. There are also several commercial E. coli strains available, as listed in Table 2, that harbor plasmids containing gene sequences encoding tRNA for rare codons, such as BL21(DE3) CodonPlus-RIL, BL21(DE3) CodonPlus-RP (Stratagene, USA), and Rosetta (DE3). Several studies have employed these excellent strategies to enhance the expression of heterologous proteins in E. coli. [13, 33, 45, 46, 71]. This approach has been successfully used to enhance the industrial production of human recombinant interferon. Co-transformation of E. coli BL21 (DE3) with the plasmid harboring human interferon-α2a gene and the argU gene that encodes for rare tRNAs for Arg (AGG/AGA) resulted in much higher levels of expression of IFN-α2a, and the recombinant protein constituted about 25% of the total proteins [26]. In another study, it was observed that production of human interferon alpha 2b protein was increased 9.5-11.5-fold by replacing the rare arginine codons with more frequently used codons [78]. The E. coli Rosetta (DE3) strain containing plasmid pRARE was used to express different human proteins, and it was shown that the yields of the recombinant protein were increased dramatically for about 35 of the 68 proteins tested [17, 75].

Table 2.Source: http://wolfson.huji.ac.il/expression/bac-strains-prot-exp.html http://www.emdmillipore.com/life-science-research/novagen http://www.invitrogen.com/1/3/stratagene-products https://www.neb.com https://www.lucigen.com https://www.genlantis.com

 

Protein Translation

Efficient translation initiation in E. coli requires a ribosomal binding site that includes the Shine-Dalgarno (SD) sequence and a translation initiation codon [69]. The Shine-Dalgarno sequence is generally located 7-9 nucleotides upstream from the initiation codon AUG [62]. It has been observed that translation initiation is more efficient from mRNAs containing the consensus SD sequence AAGGAGG. The secondary structure of ribosomal binding site is very critical for translation initiation, and the translation efficiency is further increased by the presence of a large number of thymine and adenine [35]. The efficiency of translation initiation is also affected by the nucleotide that follows the initiation codon, and it has been observed that adenine is very common in highly expressed genes [74]. Taken together, translation initiation is affected by various factors, including a consensus Shine-Dalgarno sequence, nucleotides upstream of the initiation codon, and the secondary structure of the ribosomal binding site [73]. It had also been reported that mRNA secondary structures that prevent ribosome binding might affect the protein expression to a great extent [11, 14]. It was observed that a single base change compromised the stability of the secondary structure near the SD region and resulted in a 500-fold change in the expression levels of the coat protein of RNA bacteriophage MS2. Park et al. [55, 68] have developed a very efficient and simple method for designing 5’-untranslated region (5’UTR) variants for tunable expression in E. coli. Since 5’UTR containing the SD sequence and the AU-rich sequence play a significant role in protein translation, the expression levels of recombinant proteins can be manipulated by incorporating simple variations in the 5’UTR. It has been found that secondary structures in mRNA could be disrupted by RNA helicases such as the DEAD protein of E. coli. It was shown that expression of the DEAD protein enhanced the expression of β-galactosidase by 30-fold from the T7 promoter, but there was no significant increase in the expression levels from the lac promoter [22]. However, in the absence of DEAD protein, β-galactosidase synthesis from a T7 promoter was found to be 10-fold less as compared with the expression from the lac promoter, even though the transcription rate was 10-fold higher. It was proposed that the DEAD-box protein plays an important role in stabilizing the mRNA [6, 23, 39]. The DEAD-box protein thus can be exploited to additionally improve the expression of genes with suspected problematic mRNA secondary structures.

In the bacterial genome, UAA is the most frequently used termination codon, followed by UGA and UAG. During translation, an error in reading the termination codon leads to extended protein synthesis until another termination codon is encountered in the mRNA. This read-through results in the synthesis of a larger peptide, with several additional C-terminal amino acids. It had been shown that replacing UGA with the UAA terminator in human IFN-α2b resulted in a 2-fold increase in protein expression level [66]. It had been shown that transcription terminators stabilize the mRNA by creating a stem loop structure at the 3’ end of the mRNA [49]. In E. coli, stop codon UAA is used more commonly for translation termination. The efficiency of translation termination can be improved by adding consecutive stop codons or by using a prolonged UAAU stop codon [57]. Translation errors during protein synthesis can cause frame-shift mutations, premature truncation, lower expression, and misincorporation of amino acid, and thus adversely affect the quality of recombinant protein production in E. coli [3, 56, 67]. High expression levels and correct apparent molecular mass does not always ensure the translation integrity of the recombinant protein [4, 18, 52].

 

Molecular Chaperones to Optimize Protein Folding

One major issue in the production of biopharmaceuticals using E. coli host is the accumulation of heterologous proteins, mainly as insoluble aggregates in the form of inclusion bodies. Extraction of recombinant protein from these inclusion bodies is a tedious and cumbersome process that requires denaturation and several renaturing steps to obtain a soluble and properly folded recombinant protein. One strategy to improve protein solubility in E. coli is the use of molecular chaperones. Inclusion bodies in E. coli comprise misfolded aggregates of heterologous proteins. During protein synthesis, molecular chaperones interact with nascent polypeptide chains to prevent aggregation during the folding process. Some chaperones are shown to prevent protein aggregation, while other chaperones promote refolding and solubilization of misfolded proteins [5]. The most widely used and important cytoplasmic chaperones in E. coli are DnaK, DnaJ, GrpE, GroEL, GroES, and Trigger factor (Table 3). These chaperones have been used either singly or in combination to improve the protein solubility in E. coli [1, 9, 10, 38, 64]. The most efficient chaperone combinations that are widely used to improve protein refolding are GroEL-GroES and DnaK-DnaJ-GrpE [27, 50, 84, 80]. In addition, it was observed that DnaKDnaJ-GrpE assists in the release of unfolded proteins and GroEL-GroES chaperones prevent the degradation of peptides [72]. Trigger factor has been shown to interact with GroEL and improves GroEL-substrate binding to facilitate protein folding [19]. Other chaperones such as ClpB (Hsp100), in association with DnaK, solubilizes protein and prevents aggregation [2, 47]. Other heat shock proteins, such as IpbA and IpbB, prevent aggregation of heat denatured proteins [20]. For improving the production of recombinant heterologous proteins in E. coli, different combinations of molecular chaperones should be analyzed to identify the most efficient one. It has been shown that coexpression of Skp and FkpA chaperones increased the solubility of antibody fragments in E. coli [53]. It was revealed that coexpression of GroEL–GroES resulted in production of about 65% of anti-B-type natriuretic peptide single chain antibody (scFv) in soluble form, which was 2.4-fold higher than in the absence of chaperones [40].

Table 3.ALDH3A1, aldehyde dehydrogenase 3 family, member A1; xynB, xylanase B; IL-6, Interleukin-6; Fab, fragment antigen-binding; HSP, heat shock protein.

Another strategy involves the lowering of the growth temperature to enhance the solubility of recombinant proteins [32]. Growth at a lower temperature decreases the rate of protein synthesis and thus prevents the accumulation of folding intermediates in the cytosol. Moreover, it also reduces the aggregation of protein by preventing inter- and intramolecular hydrophobic interactions, thus minimizing the formation of inclusion bodies [15]. This approach has been very efficient in enhancing the solubility of various therapeutic proteins such as interferon-α-2, human growth hormones, and Fab fragments [24, 79]. In addition, to facilitate the production of soluble recombinant proteins in E. coli, it is crucial to optimize several other parameters, such as effect of medium composition, choice of expression vectors, choice of promoters, choice of expression hosts, application of various fusion tags, rate of protein synthesis, inducer concentration, and duration of induction [16, 36, 54].

It has been suggested that the secretion of heterologous proteins into the periplasm of E. coli may provide an opportunity to produce complex therapeutic proteins. The Dsb protein family in the periplasm assists in disulfide bond formation, and the oxidizing environment in the periplasm facilitates the proper folding of heterologous proteins. Moreover, the periplasm contains very few host protein and low proteolytic activity, which facilitate the downstream processing to recover high yields of therapeutic proteins [41]. Various therapeutic proteins have been successfully manufactured by periplasmic secretion, such as commercialized Fab fragments (i.e., Leucentis and Cimza) and some full-length aglycosylated antibodies and scFvs [29, 48]. Various secretion signals have been used to target heterologous proteins to the periplasm of E. coli, such as Omp A, LamB, PhoA, STII, PelB, and endoxylanase [7]. Yim et al. [86] exploited the endoxylanase signal peptide to obtain very high-levels of expression of granulocyte colonystimulating factor (G-CSF) at 4.2 g/l in the periplasm of E. coli. Another study reported high-level periplasmic expression of insulin-like growth factor I (IGF-1), with a yield of 4.3 g/l, by using the LamB signal peptide [28]. It must be noted that besides signal peptides, the efficiency of periplasmic secretion also depends on the E. coli strain, promoter strength, and growth temperature. In some circumstances, coexpression of periplasmic chaperones, including disulfide bond oxidase (DsbA), isomerase (DsbC), and peptidylprolyl isomerase, could enhance the production of heterologous proteins in E. coli [34]. Reilly and Yansura [58] reported that overexpression of DsbA and DsbC improved the efficiency of the assembly of the light chain and the heavy chain of a full-length antibody in the periplasm, and increased the production from 0.1 to 1.05 g/l [58]. Another study reported that the yield of anti-CD20 scFv was increased, along with improved antigen binding affinity, when it was coexpressed with the periplasmic chaperone Skp [42]. Lee et al. [37] reported the development of an efficient expression system for full-length IgG in E. coli. Their study demonstrated that modification of the 5’UTR sequence and coexpression of DsbC foldase resulted in very high expression levels of heavy and light chains, and assembly of IgG was also dramatically improved in the periplasm. Under highly optimized condition, fully assembled and functionally active IgG was produced, as high as 362 mg/ml [37]. These studies clearly suggested that through proper engineering of the E. coli host, it would be possible to produce therapeutic monoclonal antibodies in E. coli in a very cost-effective and timely manner in the near future.

 

Post-Translational Modifications

E. coli is a favorite microorganism of biotechnologists for the large-scale production of therapeutic proteins [60]. However, the absence of post-translational modification processes in E. coli limits its use for the production of recombinant biopharmaceuticals. Various post-translational modifications, including glycosylation and phosphorylation, which are critical for functional activity, do not take place in E. coli due to its lack of such cellular machinery [25, 82]. N-Linked glycosylation of proteins is one of the most important post-translational modifications in eukaryotes. Wacker et al. [81] identified a novel N-linked glycosylation pathway in the bacterium Campylobacter jejuni and also showed the successful transfer of a functionally active Nglycosylation pathway into E. coli [81]. Campylobacter jejuni harbors pgl gene clusters, which are involved in the synthesis of various glycoproteins. By successfully transferring the pgl pathway into E. coli, various glycosylated proteins were produced in E. coli. Although the bacterial N-glycan structure is not similar to that seen in eukaryotes, the molecular engineering of the glycosylation pathway of C. jejuni into E. coli has paved the way for expressing glycosylated proteins in E. coli [77, 51]. Valderrama et al. [77] successfully engineered the eukaryotic glycosylation pathway in E. coli. Four eukaryotic glycosyltransferases, which include yeast uridine diphosphate-N-acetylglucosamine transferases Alg13 and Alg14, mannosyl-transferases Alg1 and Alg2, and bacterial oligosaccharyltransferase PglB from C. jejuni, were coexpressed in E. coli for synthesizing glycans, which were then successfully transferred to asparagine residues in the target eukaryotic protein [77]. This approach can also be utilized to develop glycoconjugate vaccines against several bacterial pathogens, which could be a more cost-effective and convenient alternative method to presently employed chemical-based methods of vaccine production. Currently, a glycoconjugate vaccine against Shigella dysenteriae O1, developed using this technology, has successfully cleared the phase I clinical trials. Initial efficacy and safety studies demonstrated that the glycoconjugate vaccine was safe and also elicited a strong immune response. This novel approach for glycoconjugate vaccine production using the engineered N-linked glycosylation system of Campylobacter jejuni can be exploited to produce vaccines against both gram-positive and gram-negative pathogens [8, 18, 21, 44, 83].

 

Future Perspectives

Although there are several different expression systems such as yeast, mammalian cell lines, transgenic animals, and plants that are currently being used for the production of recombinant proteins, new technological advancements are continuously being made to improve the E. coli expression system. Production of heterologous proteins in E. coli is always preferred because of the ease of genetic manipulations, well-characterized genome, accessibility of versatile plasmid vector, availability of different kinds of host strains, cost-effectiveness, and very high expression levels as compared with other expression systems. However, there are certain limitations that might hinder the efficient use of the E. coli system to overexpress heterologous genes, such as biased codon usage, protein solubility, mRNA stability, and secondary structure. The translational errors due to the presence of rare codons in heterologous proteins may result in amino acid substitutions or frame-shift mutations, which ultimately leads to undesired products. Hence, the codon usage of the recombinant protein plays a critical role in defining its expression levels. Expression of therapeutic proteins in E. coli can be enhanced by replacing rare codons in the genes with more favorable major codons. Similarly, the coexpression of genes coding for tRNAs for rare codons could increase the expression levels of therapeutic proteins in E. coli. In addition, periplasmic secretion of heterologous proteins offers several advantages, including proper folding, solubility, ease in purification, and higher yield of proteins. Antibody fragments that have been approved for therapeutic use were produced in the periplasm of E. coli, suggesting that this approach is commercially viable. Various recent studies have shown that E. coli strains can be modified specifically for each therapeutic protein to achieve high product yields as well as high-quality products.

References

  1. Arya R, Sabir JSM, Bora RS, Saini KS. 2015. Optimization of culture parameters and novel strategies to improve protein solubility. In Insoluble Proteins: Methods and Protocols. Methods in Molecular Biology series, edited by Dr. Elena Garcia Fruitos, Vol. 1258, Chapter 3. Pages 45-63; Springer Science and Business Media, New York, USA.
  2. Betiku E. 2006. Molecular chaperones involved in heterologous protein folding in Escherichia coli. Biotechnol. Mol. Biol. 1: 66-75.
  3. Brinkman U, Mattes RE, Buckel P. 1989. High-level expression of recombinant genes in Escherichia coli is dependent on the availability of the dnaY gene product. Gene 85: 109-114. https://doi.org/10.1016/0378-1119(89)90470-8
  4. Calderone TL, Stevens RD, Oas TG. 1996. High-level misincorporation of lysine for arginine at AGA codons in a fusion protein expressed in Escherichia coli. J. Mol. Biol. 262: 407-412. https://doi.org/10.1006/jmbi.1996.0524
  5. Carrio MM, Villaverde A. 2003. Role of molecular chaperones in inclusion body formation. FEBS Lett. 537: 215-221. https://doi.org/10.1016/S0014-5793(03)00126-1
  6. Chen R. 2012. Bacterial expression systems for recombinant protein production: E. coli and beyond. Biotechnol. Adv. 30: 1102-1107. https://doi.org/10.1016/j.biotechadv.2011.09.013
  7. Choi JH, Lee SY. 2004. Secretory and extracellular production of recombinant proteins using Escherichia coli. Appl. Microbiol. Biotechnol. 64: 625-635. https://doi.org/10.1007/s00253-004-1559-9
  8. Cuccui J, Wren B. 2015. Hijacking bacterial glycosylation for the production of glycoconjugates, from vaccines to humanised glycoproteins. J. Pharm. Pharmacol. 67: 338-350. https://doi.org/10.1111/jphp.12321
  9. Cui SS, Lin XZ, Shen JH. 2011. Effects of co-expression of molecular chaperones on heterologous soluble expression of the cold-active lipase Lip-948. Protein Expr. Purif. 77: 166-172. https://doi.org/10.1016/j.pep.2011.01.009
  10. de Marco A. 2007. Protocol for preparing proteins with improved solubility by co-expressing with molecular chaperones in Escherichia coli. Nat. Protoc. 2: 2632-2639. https://doi.org/10.1038/nprot.2007.400
  11. De Smit MH, van Duin, J. 1990. Secondary structure of the ribosome binding site determines translational efficiency: a quantitative analysis. Proc. Natl. Acad. Sci. USA 87: 7668-7672. https://doi.org/10.1073/pnas.87.19.7668
  12. Dieci G, Bottarelli L, Ballabeni A, Ottonello S. 2000. tRNA assisted overproduction of eukaryotic ribosomal proteins. Protein Expr. Purif. 18: 346-354. https://doi.org/10.1006/prep.2000.1203
  13. El-Baky NA, Redwan EM. 2015. Therapeutic alpha-interferons protein: structure, production, and biosimilar. Prep. Biochem. Biotechnol. 45: 109-127. https://doi.org/10.1080/10826068.2014.907175
  14. Etchegaray JP, Inouye M. 1999. Translational enhancement by an element example of molecular misreading in Alzheimer’s disease. Trends Neurosci. 21: 331-335.
  15. Fahnert B, Lilie H, Neubauer P. 2004. Inclusion bodies: formation and utilization. Adv. Biochem. Eng. Biotechnol. 89: 93-142.
  16. Feng Y, Xu Q, Yang T, Sun E, Li J, Shi D, Wu D. 2014. A novel self-cleavage system for production of soluble recombinant protein in Escherichia coli. Protein Expr. Purif. 99: 64-69. https://doi.org/10.1016/j.pep.2014.04.001
  17. Ferrer-Miralles N, Villaverde A. 2013. Bacterial cell factories for recombinant protein production; expanding the catalogue. Microb. Cell Fact. 12: 113. https://doi.org/10.1186/1475-2859-12-113
  18. Fisher AC, Haitjema CH, Guarino C, Celik E, Endicott CE, Reading CA, et al. 2011. Production of secretory and extracellular N-linked glycoproteins in Escherichia coli. Appl. Environ. Microbiol. 77: 871-881. https://doi.org/10.1128/AEM.01901-10
  19. Folwarczna J, Moravec T, Plchova H, Hoffmeisterova H, Cerovska N. 2012. Efficient expression of human papillomavirus 16 E7 oncoprotein fused to C-terminus of tobacco mosaic virus (TMV) coat protein using molecular chaperones in Escherichia coli. Protein Expr. Purif. 85: 152-157. https://doi.org/10.1016/j.pep.2012.07.008
  20. Guzzo J. 2012. Biotechnical applications of small heat shock proteins from bacteria. Int. J. Biochem. Cell Biol. 44: 1698-1705. https://doi.org/10.1016/j.biocel.2012.06.007
  21. Ihssen J, Kowarik M, Dilettoso S, Tanner C, Wacker M, Thöny-Meyer L. 2010. Production of glycoprotein vaccines in Escherichia coli. Microb. Cell Fact. 11: 9: 61. https://doi.org/10.1186/1475-2859-9-61
  22. Iost I, Dreyfus M. 1994. mRNAs can be stabilized by DEAD-box proteins. Nature 372: 193-196. https://doi.org/10.1038/372193a0
  23. Iost I, Bizebard T, Dreyfus M. 2013. Functions of DEAD-box proteins in bacteria: current knowledge and pending questions. Biochim. Biophys. Acta 1829: 866-877. https://doi.org/10.1016/j.bbagrm.2013.01.012
  24. Jensen EB, Carlsen S. 1990. Production of recombinant human growth hormone in Escherichia coli: expression of different precursors and physiological effects of glucose, acetate and salts. Biotechnol. Bioeng. 36: 1-11. https://doi.org/10.1002/bit.260360102
  25. Jenkins N. 2007. Modifications of therapeutic proteins: challenges and prospects. Cytotechnology 53: 121-125. https://doi.org/10.1007/s10616-007-9075-2
  26. Jeong W, Shin HC. 1998. Supply of the ArgU gene product allows high-level expression of recombinant human interferonalpha-2a in Escherichia coli. Biotechnol. Lett. 20: 19-22. https://doi.org/10.1023/A:1005366727366
  27. Jhamb K, Sahoo DK. 2012. Production of soluble recombinant proteins in Escherichia coli: effects of process conditions and chaperone co-expression on cell growth and production of xylanase. Bioresour. Technol. 123: 135-143. https://doi.org/10.1016/j.biortech.2012.07.011
  28. Joly JC, Leung WS, Swartz JR. 1998. Overexpression of Escherichia coli oxidoreductases increases recombinant insulinlike growth factor-I accumulation. Proc. Natl. Acad. Sci. USA 95: 2773-2777. https://doi.org/10.1073/pnas.95.6.2773
  29. Jung ST, Kang TH, Kelton W, Georgiou G. 2011. Bypassing glycosylation: engineering aglycosylated full-length IgG antibodies for human therapy. Curr. Opin. Biotechnol. 22: 858-867. https://doi.org/10.1016/j.copbio.2011.03.002
  30. Kane JF, Violand BN, Curran DF, Staten NR, Duffin KL, Bogosian G. 1992. Novel in-frame two codon translational hop during synthesis of bovine placental lactogen in a recombinant strain of Escherichia coli. Nucleic Acids Res. 20: 6707-6712. https://doi.org/10.1093/nar/20.24.6707
  31. Kane JF. 1995. Effects of rare codon clusters on high-level expression of heterologous proteins in Escherichia coli. Curr. Opin. Biotechnol. 6: 494-500. https://doi.org/10.1016/0958-1669(95)80082-4
  32. Khattar SK, Kundu PK, Gulati P, Singh V, Bughani U, Bajpaim M, Saini KS 2007. Optimization and enhanced soluble production of biologically active recombinant human p38 mitogen-activated-protein kinase (MAPK) in Escherichia coli. Protein Peptide Lett. 14: 756-760. https://doi.org/10.2174/092986607781483660
  33. Kim R, Sandler SJ, Goldman S, Yokota H, Clark AJ, Kim SH. 1998. Overexpression of archaeal proteins in Escherichia coli. Biotechnol. Lett. 20: 207-210. https://doi.org/10.1023/A:1005305330517
  34. Kolaj O, Spada S , Robin S, Wall JG. 2009. Use of folding modulators to improve heterologous protein production in Escherichia coli. Microb. Cell Fact. 8: 9. https://doi.org/10.1186/1475-2859-8-9
  35. Laursen BS, Sorensen HP, Mortensen KK, Sperling-Petersen HU. 2005. Initiation of protein synthesis in bacteria. Microbiol. Mol. Biol. Rev. 69: 101-123. https://doi.org/10.1128/MMBR.69.1.101-123.2005
  36. Lebendiker M, Danieli T. 2014. Production of prone-toaggregate proteins. FEBS Lett. 588: 236-246. https://doi.org/10.1016/j.febslet.2013.10.044
  37. Lee YJ, Lee DH, Jeong KJ. 2014. Enhanced production of human full-length immunoglobulin G1 in the periplasm of Escherichia coli. Appl. Microbiol. Biotechnol. 98: 1237-1246. https://doi.org/10.1007/s00253-013-5390-z
  38. Levy R, Weiss R, Chen G, Iverson BL, Georgiou G. 2001. Production of correctly folded Fab antibody fragment in the cytoplasm of Escherichia coli trxB gor mutants via the coexpression of molecular chaperones. Protein Expr. Purif. 23: 338-347. https://doi.org/10.1006/prep.2001.1520
  39. Linder P, Daugeron M-C. 2000. Are DEAD-box proteins becoming respectable helicases? Nat. Struct. Biol. 7: 97-99. https://doi.org/10.1038/72464
  40. Maeng BH, Nam DH, Kim YH. 2011. Coexpression of molecular chaperones to enhance functional expression of anti-BNPscFv in the cytoplasm of Escherichia coli for the detection of B-type natriuretic peptide. World J. Microbiol. Biotechnol. 27: 1391-1398. https://doi.org/10.1007/s11274-010-0590-5
  41. Makrides SC. 1996. Strategies for achieving high-level expression of genes in Escherichia coli. Microbiol. Rev. 60: 512-538.
  42. Mavrangelos C, Thiel M, Adamson PJ, Millard DJ, Nobbs S, Zola H, Nicholson IC. 2001. Increas ed yield a nd activity of soluble single-chain antibody fragments by combining highlevel expression and the Skp periplasmic chaperonin. Protein Expr. Purif. 23: 289-295. https://doi.org/10.1006/prep.2001.1506
  43. McNulty DE, Claffee BA, Huddleston MJ, Kane JF. 2003. Mistranslational errors associated with the rare arginine codon CGG in Escherichia coli. Protein Expr. Purif. 27: 365-374. https://doi.org/10.1016/S1046-5928(02)00610-1
  44. Merritt JH, Ollis AA, Fisher AC, DeLisa MP. 2013. Glycansby-design: engineering bacteria for the biosynthesis of complex glycans and glycoconjugates. Biotechnol. Bioeng. 110: 1550-1564. https://doi.org/10.1002/bit.24885
  45. Mohammed Y, El-Baky NA, Redwan NA, Redwan EM. 2012. Expression of human interferon-α8 synthetic gene under P(BAD) promoter. Biochemistry (Mosc.) 77: 1210-1219. https://doi.org/10.1134/S0006297912100136
  46. Mohammed Y, El-Bakym NA, Redwan EM. 2012. Expression, purification, and characterization of recombinant human consensus interferon-alpha in Escherichia coli under λP(L) promoter. Prep. Biochem. Biotechnol. 42: 426-447. https://doi.org/10.1080/10826068.2011.637600
  47. Nausch H, Huckauf J, Koslowski R, Meyer U, Broer I, Mikschofsky H. 2013. Recombinant production of human interleukin 6 in Escherichia coli. PLoS One 8: e54933. https://doi.org/10.1371/journal.pone.0054933
  48. Nelson AL, Reichert JM. 2009. Development trends for therapeutic antibody fragments. Nat. Biotechnol. 27: 331-337. https://doi.org/10.1038/nbt0409-331
  49. Newbury SF, Smith NH, Robinson EC, Hiles ID, Higgins CF. 1987. Stabilization of translationally active mRNA by prokaryotic REP sequences. Cell 48: 297-310. https://doi.org/10.1016/0092-8674(87)90433-8
  50. Nishihara K, Kanemori M, Kitagawa M, Yanagi H, Yura T. 1998. Chaperone coexpression plasmids: differential and synergistic roles of DnaK-DnaJ-GrpE and GroEL-GroES in assisting folding of an allergen of Japanese cedar pollen, Cryj2, in Escherichia coli. Appl. Environ. Microbiol. 64: 1694-1699.
  51. Ollis AA, Zhang S, Fisher AC, DeLisa MP. 2014. Engineered oligosaccharyltransferases with greatly relaxed acceptor-site specificity. Nat. Chem. Biol. 10: 816-822. https://doi.org/10.1038/nchembio.1609
  52. Overton TW. 2014 Recombinant protein production in bacterial hosts. Drug Discov. Today 19: 590-601. https://doi.org/10.1016/j.drudis.2013.11.008
  53. Ow DSW, Lim DYX, Nissom PM, Camattari A, Wong VVT. 2010. Co-expression of Skp and FkpA chaperones improves cell viability and alters the global expression of stress response genes during scFvD1.3 production. Microb. Cell Fact. 9: 22. https://doi.org/10.1186/1475-2859-9-22
  54. Papaneophytou CP, Kontopidis G. 2014. Statistical approaches to maximize recombinant protein expression in Escherichia coli: a general review. Protein Expr. Purif. 94: 22-32. https://doi.org/10.1016/j.pep.2013.10.016
  55. Park YS, Seo SW, Hwang S, Chu HS, Ahn JH, Kim TW, et al. 2007. Design of 5’-untranslated region variants for tunable expression in Escherichia coli. Biochem. Biophys. Res. Commun. 356: 136-141. https://doi.org/10.1016/j.bbrc.2007.02.127
  56. Parker J. 1989. Errors and alternatives in reading the universal genetics code. Microbiol. Rev. 53: 273-298.
  57. Poole ES, Brown CM, Tate WP. 1995. The identity of the base following the stop codon determines the efficiency of in vivo translational termination in Escherichia coli. EMBO J. 14: 151-158.
  58. Reilly DE, Yansura DG. 2010. Production of monoclonal antibodies in E. coli, pp295-308. In Shire SJ, Gombotz W, Bechtold-Peters K, Andya J. (eds). Current Trends in Monoclonal Antibodies Development and Manufacturing. Springer, New York.
  59. Redwan EM. 2006. Optimal gene sequence for optimal protein expression in Escherichia coli: principle requirements. Arab. J. Biotechnol. 11: 493-510.
  60. Redwan EM. 2007. Cumulative updating of approved biopharmaceuticals. Hum. Antibodies 16: 137-158.
  61. Redwan EM, Matar SM, El-Aziz GA, Serour EA. 2008. Synthesis of the human insulin gene: protein expression, scaling up and bioactivity. Prep. Biochem. Biotechnol. 38: 24-39. https://doi.org/10.1080/10826060701774312
  62. Ringquist S, Shinedling S, Barrick D, Green L, Binkley J, Stormo GD, Gold L. 1992. Translation initiation in Escherichia coli: sequences within the ribosome-binding site. Mol. Microbiol. 6: 1219-1229. https://doi.org/10.1111/j.1365-2958.1992.tb01561.x
  63. Rodriguez V, Asenjo JA, Andrews BA. 2014. Design and implementation of a high yield production system for recombinant expression of peptides. Microb. Cell Fact. 13: 65. https://doi.org/10.1186/1475-2859-13-65
  64. Ronez F, Arbault P, Guzzo J. 2012. Co-expression of the small heat shock protein, Lo18, with β-glucosidase in Escherichia coli improves solubilization and reveals various associations with overproduced heterologous protein, GroEL/ES. Biotechnol. Lett. 34: 935-939. https://doi.org/10.1007/s10529-012-0854-2
  65. Sahdev S, Khattar SK, Saini KS. 2008. Production of active eukaryotic proteins through bacterial expression systems: a review of the existing biotechnology strategies. Mol. Cell. Biochem. 307: 249-264. https://doi.org/10.1007/s11010-007-9603-6
  66. Sanchez JC, Padron G, Santana H, Herrera L. 1998. Elimination of an HuIFN alpha 2b readthrough species, produced in Escherichia coli, by replacing its natural translation stop signal. J. Biotechnol. 63: 179-186. https://doi.org/10.1016/S0168-1656(98)00073-X
  67. Seetharam R, Heeren RA, Wong EY, Braford SR, Klein BK, Aykent S, et al. 1988. Mistranslation in IGF-1 during overexpression of the protein in Escherichia coli using a synthetic gene containing low frequency codons. Biochem. Biophys. Res. Commun. 155: 518-523. https://doi.org/10.1016/S0006-291X(88)81117-3
  68. Seo SW, Yang JS, Cho HS, Yang J, Kim SC, Park JM, et al. 2014. Predictive combinatorial design of mRNA translation initiation regions for systematic optimization of gene expression levels. Sci. Rep. 4: 4515.
  69. Sørensen HP, Laursen BS, Mortensen KK, Sperling-Petersen HU. 2002. Bacterial translation initiation — mechanism and regulation. Recent Res. Dev. Biophys. Biochem. 2: 243-270.
  70. Sørensen HP, Sperling-Petersen HU, Mortensen KK. 2003. A favorable solubility partner for the recombinant expression of streptavidin. Protein Expr. Purif. 32: 252-259. https://doi.org/10.1016/j.pep.2003.07.001
  71. Sørensen HP, Sperling-Petersen HU, Mortensen KK. 2003. Dialysis strategies for protein refolding: preparative streptavidin production. Protein Expr. Purif. 31: 149-154. https://doi.org/10.1016/S1046-5928(03)00133-5
  72. Sorensen HP, Mortensen KK. 2005. Soluble expression of recombinant proteins in the cytoplasm of Escherichia coli. Microb. Cell Fact. 4: 1. https://doi.org/10.1186/1475-2859-4-1
  73. Sprengart ML, Porter AG. 1997. Functional importance of RNA interactions in selection of translation initiation codons. Mol. Microbiol. 24: 19-28. https://doi.org/10.1046/j.1365-2958.1997.3161684.x
  74. Stenstrom C, Jin H, Major L, Tate W, Isaksson LA. 2001. Codon bias at the 3’-side of the initiation codon is correlated with translation initiation efficiency in Escherichia coli. Gene 263: 273-284. https://doi.org/10.1016/S0378-1119(00)00550-3
  75. Tegel H, Tourle S, Ottosson J, Persson A. 2010. Increased levels of recombinant human proteins with the Escherichia coli strain Rosetta (DE3). Protein Expr. Purif. 69: 159-167. https://doi.org/10.1016/j.pep.2009.08.017
  76. Terra VS, Mills DC, Yates LE, Abouelhadid S, Cuccui J, Wren BW. 2012. Recent developments in bacterial protein glycan coupling technology and glycoconjugate vaccine design. J. Med. Microbiol. 61: 919-926. https://doi.org/10.1099/jmm.0.039438-0
  77. Valderrama-Rincon JD, Fisher AC, Merritt JH, Fan YY, Reading CA, Chhiba K, et al. 2012. An engineered eukaryotic protein glycosylation pathway in Escherichia coli. Nat. Chem. Biol. 8: 434-436. https://doi.org/10.1038/nchembio.921
  78. Valente CA, Prazeres DMF, Cabral JMS, Monteriro GA. 2004. Translation feature of human alpha 2b interferon production in Escherichia coli. Appl. Environ. Microbiol. 70: 5033-5036. https://doi.org/10.1128/AEM.70.8.5033-5036.2004
  79. Vasina JA, Baneyx F. 1997. Expression of aggregation prone recombinant proteins at low temperatures: a comparative study of the Escherichia coli cspA and tac promoters systems. Protein Expr. Purif. 9: 211-218. https://doi.org/10.1006/prep.1996.0678
  80. Voulgaridou GP, Mantso T, Chlichlia K, Panayiotidis MI, Pappa A. 2013. Efficient E. coli expression strategies for production of soluble human crystallin ALDH3A1. Plos One 8: e56582. https://doi.org/10.1371/journal.pone.0056582
  81. Wacker M, Linton D, Hitchen PG, Nita-Lazar M, Haslam SM, North SJ, et al. 2002. N-linked glycosylation in Campylobacter jejuni and its functional transfer into E. coli. Science 298: 1790-1793. https://doi.org/10.1126/science.298.5599.1790
  82. Walsh G, Jefferis R. 2006. Post-translational modifications in the context of therapeutic proteins. Nat. Biotechnol. 24: 1241-1252. https://doi.org/10.1038/nbt1252
  83. Wetter M, Kowarik M, Steffen M, Carranza P, Corradin G, Wacker M. 2013. Engineering, conjugation, and immunogenicity assessment of Escherichia coli O121 O antigen for its potential use as a typhoid vaccine component. Glycoconj. J. 30: 511-522. https://doi.org/10.1007/s10719-012-9451-9
  84. Yan X, Hu S, Guan YX, Yao SJ. 2012. Coexpres sion of chaperonin GroEL/GroES markedly enhanced soluble and functional expression of recombinant human interferongamma in Escherichia coli. Appl. Microbiol. Biotechnol. 93: 1065-1074. https://doi.org/10.1007/s00253-011-3599-2
  85. Yarian C, Marszalek M, Sochacka E, Malkiewicz A, Guenther R, Miskiewicz A, Agris PF. 2000. Modified nucleoside dependent Watson-Crick and wobble codon binding by tRNALysUUU species. Biochemistry 39: 13390- 13395. https://doi.org/10.1021/bi001302g
  86. Yim S, Jeong K, Chang H, Lee S. 2001. High-level secretory production of human granulocytes-colony stimulating factor by fed-batch culture of recombinant Escherichia coli. Bioprocess Biosyst. Eng. 24: 249-254. https://doi.org/10.1007/s004490100267

Cited by

  1. Constitutive expression of the sRNA GadY decreases acetate production and improves E. coli growth vol.14, pp.None, 2015, https://doi.org/10.1186/s12934-015-0334-1
  2. Production and Purification of Recombinant Proteins from Escherichia coli vol.3, pp.3, 2015, https://doi.org/10.1002/cben.201600002
  3. The E. coli pET expression system revisited—mechanistic correlation between glucose and lactose uptake vol.100, pp.20, 2016, https://doi.org/10.1007/s00253-016-7620-7
  4. Expression and Purification of C-Peptide Containing Insulin Using Pichia pastoris Expression System vol.2016, pp.None, 2016, https://doi.org/10.1155/2016/3423685
  5. Hib Vaccines: Past, Present, and Future Perspectives vol.2016, pp.None, 2016, https://doi.org/10.1155/2016/7203587
  6. Rapid label‐free quantitative analysis of the E. coli BL21(DE3) inner membrane proteome vol.16, pp.1, 2016, https://doi.org/10.1002/pmic.201500304
  7. Protein folding in the cell envelope of Escherichia coli vol.1, pp.8, 2015, https://doi.org/10.1038/nmicrobiol.2016.107
  8. Recombinant production of medium- to large-sized peptides in Escherichia coli using a cleavable self-aggregating tag vol.15, pp.None, 2016, https://doi.org/10.1186/s12934-016-0534-3
  9. Synthetic Microbial Ecology: Engineering Habitats for Modular Consortia vol.8, pp.None, 2017, https://doi.org/10.3389/fmicb.2017.01125
  10. Construction of pDUO: A bicistronic shuttle vector series for dual expression of recombinant proteins vol.89, pp.None, 2015, https://doi.org/10.1016/j.plasmid.2016.12.001
  11. Simple monitoring of cell leakiness and viability in Escherichia coli bioprocesses—A case study vol.17, pp.6, 2015, https://doi.org/10.1002/elsc.201600204
  12. 재조합 대장균의 고농도 배양과 유도조건 최적화를 통한 Bacillus 유래 esterase의 생산 vol.45, pp.2, 2017, https://doi.org/10.4014/mbl.1703.03002
  13. Synthetic Nanochaperones Facilitate Refolding of Denatured Proteins vol.11, pp.10, 2015, https://doi.org/10.1021/acsnano.7b05947
  14. Improvement in the production of the human recombinant enzyme N-acetylgalactosamine-6-sulfatase (rhGALNS) in Escherichia coli using synthetic biology approaches vol.7, pp.None, 2015, https://doi.org/10.1038/s41598-017-06367-w
  15. A combination of HPLC and automated data analysis for monitoring the efficiency of high-pressure homogenization vol.16, pp.None, 2017, https://doi.org/10.1186/s12934-017-0749-y
  16. Improving E. coli growth performance by manipulating small RNA expression vol.16, pp.None, 2015, https://doi.org/10.1186/s12934-017-0810-x
  17. Generation of a platform strain for ionic liquid tolerance using adaptive laboratory evolution vol.16, pp.None, 2015, https://doi.org/10.1186/s12934-017-0819-1
  18. Intensification of Organophosphorus Hydrolase Synthesis by Using Substances with Gas-Transport Function vol.7, pp.12, 2015, https://doi.org/10.3390/app7121305
  19. Microbial production of toxins from the scorpion venom: properties and applications vol.102, pp.15, 2015, https://doi.org/10.1007/s00253-018-9122-2
  20. New trends in aggregating tags for therapeutic protein purification vol.40, pp.5, 2015, https://doi.org/10.1007/s10529-018-2543-2
  21. Production of Ready-To-Use Functionalized Sup35 Nanofibrils Secreted by Komagataella pastoris vol.12, pp.9, 2015, https://doi.org/10.1021/acsnano.8b04450
  22. Hirudin variants production by genetic engineered microbial factory vol.34, pp.2, 2015, https://doi.org/10.1080/02648725.2018.1506898
  23. Recent Developments in Recombinant Protein–Based Dengue Vaccines vol.9, pp.None, 2015, https://doi.org/10.3389/fimmu.2018.01919
  24. Impact of Glycerol as Carbon Source onto Specific Sugar and Inducer Uptake Rates and Inclusion Body Productivity in E. coli BL21 ( DE3 ) vol.5, pp.1, 2015, https://doi.org/10.3390/bioengineering5010001
  25. A Direct Droplet Digital PCR Method for E. coli Host Residual DNA Quantification vol.9, pp.4, 2018, https://doi.org/10.4236/pp.2018.94009
  26. Progress in biopharmaceutical development vol.65, pp.3, 2015, https://doi.org/10.1002/bab.1617
  27. Higher‐order structure and conformational change in biopharmaceuticals vol.93, pp.9, 2018, https://doi.org/10.1002/jctb.5682
  28. Leptospira interrogans thermolysin refolded at high pressure and alkaline pH displays proteolytic activity against complement C3 vol.19, pp.None, 2015, https://doi.org/10.1016/j.btre.2018.e00266
  29. The production of a recombinant tandem single chain fragment variable capable of binding prolamins triggering celiac disease vol.18, pp.None, 2015, https://doi.org/10.1186/s12896-018-0443-0
  30. Custom made inclusion bodies: impact of classical process parameters and physiological parameters on inclusion body quality attributes vol.17, pp.None, 2015, https://doi.org/10.1186/s12934-018-0997-5
  31. E. coli HMS174(DE3) is a sustainable alternative to BL21(DE3) vol.17, pp.None, 2015, https://doi.org/10.1186/s12934-018-1016-6
  32. Inclusion Body Bead Size in E. coli Controlled by Physiological Feeding vol.6, pp.4, 2015, https://doi.org/10.3390/microorganisms6040116
  33. Chasing bacterial chassis for metabolic engineering: a perspective review from classical to non‐traditional microorganisms vol.12, pp.1, 2015, https://doi.org/10.1111/1751-7915.13292
  34. Recent Developments in Bioprocessing of Recombinant Proteins: Expression Hosts and Process Development vol.7, pp.None, 2015, https://doi.org/10.3389/fbioe.2019.00420
  35. Platinum and Palladium Bio-Synthesized Nanoparticles as Sustainable Fuel Cell Catalysts vol.7, pp.None, 2015, https://doi.org/10.3389/fenrg.2019.00066
  36. Whole synthetic pathway engineering of recombinant protein production vol.116, pp.2, 2015, https://doi.org/10.1002/bit.26855
  37. Perspectives of inclusion bodies for bio-based products: curse or blessing? vol.103, pp.3, 2015, https://doi.org/10.1007/s00253-018-9569-1
  38. Soluble expression of IGF1 fused to DsbA in SHuffle™ T7 strain: optimization of expression and purification by Box-Behnken design vol.103, pp.8, 2019, https://doi.org/10.1007/s00253-019-09719-w
  39. Recombinant production of ESAT-6 antigen in thermoinducible Escherichia coli: the role of culture scale and temperature on metabolic response, expression of chaperones, and architecture of inclusion b vol.24, pp.4, 2019, https://doi.org/10.1007/s12192-019-01006-x
  40. New tools for recombinant protein production in Escherichia coli : A 5‐year update vol.28, pp.8, 2019, https://doi.org/10.1002/pro.3668
  41. Heterologous expression and biochemical characterisation of cyanotoxin biosynthesis pathways vol.36, pp.8, 2015, https://doi.org/10.1039/c8np00063h
  42. Versatile biomanufacturing through stimulus-responsive cell-material feedback vol.15, pp.10, 2015, https://doi.org/10.1038/s41589-019-0357-8
  43. Elastin-like polypeptide fusions for high-level expression and purification of human IFN-γ in Escherichia coli vol.585, pp.None, 2015, https://doi.org/10.1016/j.ab.2019.113401
  44. Engineering the flagellar type III secretion system: improving capacity for secretion of recombinant protein vol.18, pp.None, 2015, https://doi.org/10.1186/s12934-019-1058-4
  45. Generation of an E. coli platform strain for improved sucrose utilization using adaptive laboratory evolution vol.18, pp.None, 2015, https://doi.org/10.1186/s12934-019-1165-2
  46. The Lazarus Escherichia coli Effect: Recovery of Productivity on Glycerol/Lactose Mixed Feed in Continuous Biomanufacturing vol.8, pp.None, 2015, https://doi.org/10.3389/fbioe.2020.00993
  47. Repetitive Fed-Batch: A Promising Process Mode for Biomanufacturing With E. coli vol.8, pp.None, 2015, https://doi.org/10.3389/fbioe.2020.573607
  48. Fusion tags to enhance heterologous protein expression vol.104, pp.6, 2015, https://doi.org/10.1007/s00253-020-10402-8
  49. Perspectives for Glyco-Engineering of Recombinant Biopharmaceuticals from Microalgae vol.9, pp.3, 2015, https://doi.org/10.3390/cells9030633
  50. ASA 3 P: An automatic and scalable pipeline for the assembly, annotation and higher-level analysis of closely related bacterial isolates vol.16, pp.3, 2015, https://doi.org/10.1371/journal.pcbi.1007134
  51. A lipoprotein lipase–GPI-anchored high-density lipoprotein–binding protein 1 fusion lowers triglycerides in mice: Implications for managing familial chylomicronemia syndrome vol.295, pp.10, 2020, https://doi.org/10.1074/jbc.ra119.011079
  52. Multiphase matrix of silica, culture medium and air for 3D mammalian cell culture vol.72, pp.2, 2015, https://doi.org/10.1007/s10616-020-00376-w
  53. A Heterologous Viral Protein Scaffold for Chimeric Antigen Design: An Example PCV2 Virus Vaccine Candidate vol.12, pp.4, 2015, https://doi.org/10.3390/v12040385
  54. Multiplex Generation, Tracking, and Functional Screening of Substitution Mutants Using a CRISPR/Retron System vol.9, pp.5, 2020, https://doi.org/10.1021/acssynbio.0c00002
  55. Cost analysis based on bioreactor cultivation conditions: Production of a soluble recombinant protein using Escherichia coli BL21(DE3) vol.26, pp.None, 2015, https://doi.org/10.1016/j.btre.2020.e00441
  56. Production of Active Recombinant Hyaluronidase Inclusion Bodies from Apis mellifera in E. coli Bl21(DE3) and characterization by FT-IR Spectroscopy vol.21, pp.11, 2015, https://doi.org/10.3390/ijms21113881
  57. Synthetic Glycobiology: Parts, Systems, and Applications vol.9, pp.7, 2015, https://doi.org/10.1021/acssynbio.0c00210
  58. Preparation and Characterization of β-glucosidase Films for Stabilization and Handling in Dry Configurations vol.21, pp.8, 2020, https://doi.org/10.2174/1389201020666191202145351
  59. Aggregation-prone peptides modulate activity of bovine interferon gamma released from naturally occurring protein nanoparticles vol.57, pp.None, 2015, https://doi.org/10.1016/j.nbt.2020.02.001
  60. Current technologies to endotoxin detection and removal for biopharmaceutical purification vol.117, pp.8, 2020, https://doi.org/10.1002/bit.27362
  61. Experimental Evaluation of In Silico Selected Signal Peptides for Secretory Expression of Erwinia Asparaginase in Escherichia coli vol.26, pp.3, 2015, https://doi.org/10.1007/s10989-019-09961-w
  62. Aspergillus: A Powerful Protein Production Platform vol.10, pp.9, 2015, https://doi.org/10.3390/catal10091064
  63. High pressure homogenization is a key unit operation in inclusion body processing vol.7, pp.None, 2015, https://doi.org/10.1016/j.btecx.2020.100022
  64. CRISPR base editing and prime editing: DSB and template-free editing systems for bacteria and plants vol.5, pp.4, 2015, https://doi.org/10.1016/j.synbio.2020.08.003
  65. Global dynamics of the chemostat with overflow metabolism vol.82, pp.3, 2015, https://doi.org/10.1007/s00285-021-01566-6
  66. Hydrofoam and oxygen headspace bioreactors improve cell‐free therapeutic protein production yields through enhanced oxygen transport vol.37, pp.2, 2015, https://doi.org/10.1002/btpr.3079
  67. Synthetic Biological Approaches for Optogenetics and Tools for Transcriptional Light‐Control in Bacteria vol.5, pp.5, 2015, https://doi.org/10.1002/adbi.202000256
  68. Consolidated Bioprocessing: Synthetic Biology Routes to Fuels and Fine Chemicals vol.9, pp.5, 2015, https://doi.org/10.3390/microorganisms9051079
  69. Soluble production of a full-length human papillomavirus type 16 L1 protein by Escherichia coli vol.6, pp.2, 2015, https://doi.org/10.21931/rb/2021.06.02.4
  70. At-Line Reversed Phase Liquid Chromatography for In-Process Monitoring of Inclusion Body Solubilization vol.8, pp.6, 2015, https://doi.org/10.3390/bioengineering8060078
  71. Strategies for the Production of Soluble Interferon-Alpha Consensus and Potential Application in Arboviruses and SARS-CoV-2 vol.11, pp.6, 2021, https://doi.org/10.3390/life11060460
  72. Study of Anti-Inflammatory and Analgesic Activity of Scorpion Toxins DKK-SP1/2 from Scorpion Buthus martensii Karsch (BmK) vol.13, pp.7, 2021, https://doi.org/10.3390/toxins13070498
  73. Engineering of a robust Escherichia coli chassis and exploitation for large-scale production processes vol.67, pp.None, 2015, https://doi.org/10.1016/j.ymben.2021.05.011
  74. Microbial synthesis of wax esters vol.67, pp.None, 2021, https://doi.org/10.1016/j.ymben.2021.08.002
  75. COVID-19 vaccine platforms: Delivering on a promise? vol.17, pp.9, 2015, https://doi.org/10.1080/21645515.2021.1911204
  76. Broad-spectrum nanoparticles against bacteriophage infections vol.13, pp.44, 2021, https://doi.org/10.1039/d1nr04936d
  77. Cloning and expression of antibody fragment (Fab) I: Effect of expression construct and induction strategies on light and heavy chain gene expression vol.176, pp.None, 2015, https://doi.org/10.1016/j.bej.2021.108189
  78. Combining Ancestral Reconstruction with Folding-Landscape Simulations to Engineer Heterologous Protein Expression vol.433, pp.24, 2021, https://doi.org/10.1016/j.jmb.2021.167321
  79. Cascaded processing enables continuous upstream processing with E. coli BL21(DE3) vol.11, pp.1, 2015, https://doi.org/10.1038/s41598-021-90899-9