DOI QR코드

DOI QR Code

Improving Protein Production on the Level of Regulation of both Expression and Secretion Pathways in Bacillus subtilis

  • Song, Yafeng (Tianjin Institute of Industrial Biotechnology, Chinese Academy of Sciences) ;
  • Nikoloff, Jonas M. (Tianjin Institute of Industrial Biotechnology, Chinese Academy of Sciences) ;
  • Zhan, Dawei (Tianjin Institute of Industrial Biotechnology, Chinese Academy of Sciences)
  • Received : 2015.01.12
  • Accepted : 2015.02.13
  • Published : 2015.07.28

Abstract

The well-characterized gram-positive bacterium Bacillus subtilis is an outstanding industrial candidate for protein expression owing to its single membrane and high capacity of secretion, simplifying the downstream processing of secretory proteins. During the last few years, there has been continuous progress in the illustration of secretion mechanisms and application of this robust host in various fields of life science, such as enzyme production, feed additives, and food and pharmaceutical industries. Here, we review the developments of Bacillus subtilis as a highly promising expression system illuminating strong chemical- and temperatureinducible and other types of promoters, strategies for ribosome-binding-site utilization, and the novel approach of signal peptide selection. Furthermore, we outline the main steps of the Sec pathway and the relevant elements as well as their interactions. In addition, we introduce the latest discoveries of Tat-related complex structures and functions and the countless applications of this full-folded protein secretion pathway. This review also lists some of the current understandings of ATP-binding cassette transporters. According to the extensive knowledge on the genetic modification strategies and molecular biology of Bacillus subtilis, we propose some suggestions and strategies for improving the yield of intended productions. We expect this to promote striking future developments in the optimization and application of this bacterium.

Keywords

Introduction

The gram-positive bacterium Bacillus subtilis is by far one of the most widely used hosts for the mass production of recombinant proteins in biotechnology. The popularity of this remarkable microorganism, in both basic research and practical applications of secreting target proteins, is based on its characteristics, such as its status of being generally recognized as safe (GRAS), being sequenced totally, ease of manipulation, absence of significant codon bias, and genetically well-characterized expression systems. Furthermore, its production has widely been used in fields of red and white biotechnology, such as the pharmaceutical, food, and feed industries [21, 23, 64, 115, 132].

In general, Bacillus subtilis is much more suitable for the production of secretory recombinant proteins than Escherichia coli. Lacking an outer membrane, B. subtilis easily targets proteins of interest to the culture medium, simplifying downstream processing and eliminating the need of cell lysis and chemical process techniques for the isolation and purification of overexpressed products. Well-characterized genetics and the availability of an increasing number of cloning vectors and mutant host strains are provided by the Bacillus Genetic Stock Centre (BGSC) [99]. Being a long journey for the cargo proteins from birth to maturity, the steps from transcription to secretion include a variety of unknown bottlenecks for the final production [61, 63, 79, 90].

During several decades, we have witnessed the advancements of different industrial workhorses in aspects of genetic engineering and secretory machine mechanisms. Shedding light upon the latest improvements of this expression system, this review concentrates on current understandings of the Sec and Tat pathways, transformation methods, and genetic engineering strategies. This review will drive developments of Bacillus subtilis as a high-performance candidate in the production of secretory recombinant proteins.

 

Expression System

To improve the protein production, regulation can take place at the levels of both transcription and translation. Up to now, the basic economical and effective approach still remains the optimization of promoters at the level of transcription. Therefore, rational choice of the adequate promoter system for a specific protein of interest remains important. However, selecting suitable ribosome-binding sites (RBS) and signal peptides (SP) has also led to high yields of protein production, but at the level of translation.

Promoter

Regulation of gene expression in Bacillus subtilis primarily occurs during transcription, in which RNA polymerase (RNAP) binds to the corresponding promoter sequences. This RNAP consists of the core enzyme (α, α’, β, β’) and the σ-factor, which determines the specifity of RNAPs. The σ- factors of vegetative cells include σA, σB, σC, σD, σH, σL and σE, σF, σG, σK, where the latter group belongs to the sporulation-specific σ-factors [30]. In addition, the use of pro-protein sequences and anti-σ-factor proteins, in order to modulate σ-factor activity, indicates that proteins are expressed under sophisticated conditions of transcriptional control.

Promoter regions share several characteristics [32]: the consensus -35 sequence of TTGACA, a -10 sequence of TATA AT, a nd a spacer between t he - 35 a nd -10 regions usually comprising 17 base pairs, with some comprising 18 base pairs [73]. A -15 region (or TGn motif) from position -17 to -14 contains TRTG (where R stands for A or G). purine initiation principle means that the translational initiation site usually includes A or G. Most promoters having 4-7 base pairs between the -10 and +1 sites usually are enriched by AT pairs. Another AT-rich region upstream from the -35 region and designated as UP element also counts as important in the efficient application of Bacillus subtilis promoters. It can be considered as a transcription-enhancing region interacting with the C-terminal domain (CTD) of the α-subunit of the RNA polymerase [1].

It was demonstrated that enrichment of AT in the natural UP element of PgroE leads to a 2-fold increased activity of the report gene bgaB [86]. There is distant genetic controlling in the yqiHIK promoter [24]: The activity and osmotic regulation of the yqiHIK promoter are also influenced by a 53 bp sequence comprising an AT-rich DNA segment, positioned 180 bp upstream of the -35 sequence. When changing the core region of PgroE into a consensus -35 and -10 sequence, the promoter activity increases about 9-fold and 16-fold, respectively, and 100- fold when both changes are combined in one single promoter. Point mutations within the -15 region also result in significantly increased activity. After combining the optimized sequences of the UP element and the -35, -15, and -10 regions, intracellular product accumulation of the reporter gene bgaB increases from 9% to 30% of the total cellular wild-type protein amount. Thus on one hand, it is useful to link all these optimized elements in one single promoter. On the other hand, one needs to deal with the degradation of overexpressed recombinant proteins by knocking out protease genes, since efficient transcription does not guarantee a concomitant increase in translation and correctly folded state of proteins.

During the past few years and according to these major elements, there have been many reports about strong promoters, their corresponding expression systems, and other types of promoters in order to facilitate the experimental and industrial utilization. Here, we summarize as follows (Table 1).

Table 1.Different types of promoters and corresponding expression systems of Bacillus subtilis.

Apart from searching for suitable promoters for the expression of specific proteins in published literature or patents, promoter screening via the pHT06 vector allows the cloning and analysis of strong promoters in Bacillus subtilis on X-Gal plates and IPTG added in appropriate concentrations [85].

Ribosome-Binding Site

The RBS element plays an important role in the translational efficiency of Bacillus subtilis even though the factors determining mRNA stability are far from being completely identified and understood [6]. It was demonstrated that the important elements involved are the Shine-Dalgarno (SD) sequence, known to form Watson-Crick base pairs with the 16S rRNA of the 30S subunit, the spacer length between the SD sequence and initiation codon, the initiation codon itself and the spatial organization of the initiation domain of the mRNA [108].

It has been reported that a strong optimal RBS appears to be crucial for increased stability of the σB-dependent general stress protein (gsiB) mRNA [43]. This is consistent with the observations showing that the mRNA half-life is rapidly reduced after removing a stem-loop structure at the 5’ end or diminishing the strength of the RBS [31]. Furthermore, inactivation of the RBS or the start codon results in a 4-fold decrease of the ermC mRNA half-life. This plasmid-borne erythromycin resistance (Emr) gene is a 254-nucleotide model mRNA used to study the effects of translational signals and ribosome transit on mRNA decay in Bacillus subtilis [97]. Additionally, the sequence upstream of the translational start codon is crucial for the long halflife of the gsiB mRNA rather than the translated sequence, which is revealed by the northern blotting experiments with different gsiB-lacZ fusions.

To increase mRNA stability in order to elevate the production of the target protein, Phan et al. [87] proposed a 5’-mRNA controlled stabilizing element (CoSE) by combining a transcriptional operator lacO stem-loop with a strong RBSgsiB, fused to the promoter groE. With further suitable spacer in between, the mRNA stability presented a half-life of more than 60 min. This indicates that this method is a promising tool for enhancing protein expression levels.

Moreover, for artificial elements, tools for automated design of synthetic RBSs have been developed to control protein expression [93, 94, 112]. In their statistical thermodynamic equilibrium model, the translation initiation rate is quantified according to the strength of the molecular interactions between an mRNA transcript and the 30S ribosome complex. The attack of 5’ endoribonucleases can be prevented when the initiating ribosomes mask the nuclease recognition sites near the RBS; thus, mRNA stability is tightly coupled to the translation initiation frequency [14]. The fact that the spacer length between a SD sequence and the start codon influences the translational efficiency, combined with the statistical thermodynamic equilibrium model, led to the development of the software RBSDesigner. This software roughly enables to define yield of protein expression by specific RBS design [77].

Signal Peptide

Proteins destined to be transferred across the cytoplasmic membrane possess N-terminal signal peptides, which contain three distinct regions: a positively charged Nterminal domain, followed by a hydrophobic core region, the so-called H-region, and the C-terminal region with a signal peptidase cleavage site [90] (Fig. 1). Signal peptides consist of short stretches of amino acids and will be immediately removed and degraded by specialized signal peptidases and signal peptide peptidases (SPPases) after the protein delivery to the correct destination. During translocation, the positively charged N-terminal domain interacts with both the translocation machinery and negatively charged phospholipids in the lipid bilayer of the membrane. Helix-breaking residues of glycine or proline, which are frequently present in the middle of H-regions, allow the signal peptide to form a hairpin-like structure that can insert into the membrane.

Fig. 1.S cheme of t he s ignal p eptides of different s ecreting pathways of Bacillus subtilis. (A), (B), (C), and (D) represents the Sec-dependent signal peptide, the Tat-dependent signal peptide, the ABC-dependent signal peptide, and the Com-dependent signal peptide, respectively. The arrowhead points to the cleavage site.

Respectively, signal peptides can be divided into two types, depending on the signal peptidases that are responsible for the release of mature proteins to final positions by cleavage of signal peptides, as described by Pohl and Harwood [90].

Nevertheless, signal peptides can also be assigned to other classes, according to export pathways and signal peptidase cleavage sites, as summarized by Tjalsma et al. [102, 103] and Ling et al. [63]. Both papers report of Sectype signal peptides and twin-arginine signal peptides. The main difference is the type of SP according to their signal peptidase. The lantibiotic and pheromone signal peptides consist of only N- and C-terminal domains, but lack the Hdomain, which is removed by specific ABC transporters.

Furthermore, the Com-dependent pseudolipin-like signal peptides, detected at four proteins of Bacillus subtilis, are designated as ComGC/GD/GE and GG. Owning 33 residues on average, its consensus sequence (K-G-F) is located between the N and H-domains. Bypassing Sec and Tat pathways by its own specific pathway, pseudolipin-like signal peptidases work at the cytoplasmatic side of the membrane. By this, they take part in minomethylation and DNA-binding processes to the membrane during transformation [16, 102].

The other type of signal peptide is followed by a propeptide. In this case, the signal peptide is removed by a Type I enzyme and the propeptide is removed either autocatalytically or by a coexisting protease.

As the hydrophobicity of the signal peptide is a critical factor for early stage protein secretion [125], researchers have developed the signal peptide prediction tool SignalP, which helps choosing signal peptides. It treats signal peptides with a calculated high D-score (usually above 0.5) as the suitable ones to enhance the production. Unfortunately, the D-score and its real secretion amounts do not seem to correlate, as was determined by a complex pattern of events. In order to construct a powerful strategy in optimizing the secretion of heterologous proteins, Brockmeier et al. [11] reported a SP library to screen the individually optimal signal peptides for each secreted heterologous protein, which was achieved by a high-throughput screening method. Cutinase, an enzyme from Fusarium solani pisi and the metagenomic cytoplasmatic esterase, which acts as a model reporter for naturally unsecreted proteins, were chosen as two lipolytic enzyme reporters of heterologous protein secretion. A total of 148 of the 173 SP-cutinase fusions could be expressed in Bacillus subtilis, with different lipolytic activity of the culture supernatants. Seven SPs with strong, moderate, and low efficiencies were further analyzed by western blotting and pulse-chase experiments. The results indicate that translocation and SP processing efficiencies are not the sole determinants responsible for protein secretion. When the most efficient SP for cutinase is fused to the esterase, the secretion level only amounts to about 5% of the most efficient SP for esterase. This phenomenon underlines the secretion target specificity of the best SP, as the interaction between the SP and the mature protein is influenced by the composition of the amino acids of the two fragments.

Consistent results had been reported by Degering et al. [18], who constructed a signal peptide library containing 173 predicted SPs originating from Bacillus subtilis and 220 from the Bacillus licheniformis. They identified two SPs resulting in a 7-fold increase of subtilisin activity in the medium supernatant, compared with wild-type SP. These two SPs are not from high secreted enzymes like AprE and NprE. This indicates that SPs from highly secreted proteins do not necessarily lead to efficient secretion of proteins. They also recommend using Bacillus subtilis for preliminary high-throughput screenings of SP libraries of specific proteins. After this, one can transfer the best-performing SP to a production strain, in case the production strain, such as Bacillus licheniformis, is not amenable to molecular manipulations, including DNA transformation.

Another alternative for increasing cutinase secretion by finding the beneficial amino acids of the SP in Bacillus subtilis is saturation mutagenesis of the positively charged N-domain (positions 2–7) [12]. This method enables the identification of single point mutants leading to 2- to 4-fold increased amounts of cutinase expression. Not only is the amount of cutinase in the supernatant improved, but also the quality of the protein with respect to its specific activity, even though a reduction of the net charge of the SP from +3 to +2 is observed, which usually is associated with a less effective interaction of the SP with its binding partners [84].

Owing to our limited knowledge about the secretion mechanism, it is still difficult to significantly improve protein secretion by rational design of amino acids. Both the construction of SP libraries and directed evolution, and their combination, seem to be efficient approaches. However, it should be noted that an appropriate high-throughput screening method for specific proteins is necessary at the same time.

 

Process of Secretion Pathways and Related Proteins

Proteins are transported to their destinations after synthesis through distinct mechanisms [115] (Fig. 2). Bacillus subtilis has evolved several pathways to fulfill this request in order to adapt to the changing environment: the general protein secretion pathway (Sec), the two-arginine pathway (Tat), and the ATP-binding cassette (ABC) transporters. Here, we mainly introduce the detail of the Sec pathway and the relevant proteins in this channel. We also focus on the latest discoveries about the Tat pathway. The ABC transporters are simply described.

Fig. 2.Scheme of the different secreting pathways of Bacillus subtilis. (A) The Sec pathway: the signal recognition particle (SRP) recognizes the signal peptides of the ribosome-nascent chain complexes, and the cytoplasmic chaperones guarantee the translocation-competent state. Then pre-proteins combine with the SRP receptor FtsY (Aa), or translocate via the SecYEG translocase complex (Ab). Then the signal peptide is cut and degraded by the Spases, SppA and TepA, respectively, and finally modified and checked by other complexes, such as BdbCD and HtrAB. (B) The Tat pathway: pre-proteins fold in the cytoplasm with the help of cofactors, and then pass through the Tat translocase under the presence of the pH gradient across the cytosolic membrane. After translocation, type I Spases cut down the signal peptides and the remaining proteins secrete into the medium. (C) The ABC pathway: proteins secrete through the ATP-binding cassette, by mechanisms that have not been clearly understood yet. (D) Misfolded proteins are degraded by ClpCP, ClpXP, and ClpEP in the cytoplasm or by extracytoplasmatic proteases such as NprB, AprE, Epr, and/or other proteases.

Sec Pathway

Recognition and targeting. Immediately after the synthesis of precursor proteins, the cytoplasmic chaperones guarantee that preproteins gather the state of translocationcompetence before their signal peptides are recognized by the signal recognition particle (SRP) and are combined with the SRP receptor FtsY [2, 127].

The highly conserved ribonucleoprotein complex SRP consists of a scRNA, two histone-like proteins (HBsu), and Ffh, which is a 446-amino-acid protein interacting with scRNA and forming a stable complex. The SRP RNA (long SRP RNA) of Bacillus subtilis contains an Alu domain, which arrests protein biosynthesis by competing with elongation factors that bind on ribosomes. Retarding translation gives SRP time to interact with its membrane receptor FtsY. The DNA-binding protein HU1, a part of HBsu, has been suggested to be part of the bacterial Alu domain [48].

Both Ffh and FtsY belong to the widely conserved family of SRP-GTPase. The GTP hydrolysis at FtsY and SRP affects the release of the ribosome-nascent chain complex (RNC), the complex comprising the nascent peptide chain, tRNA and the ribosome, from SRP to the Sec translocation channel [116]. However, a third SRP-GTPase, named FlhF (flagellum-associated protein), seems to be dispensable for protein secretion. It has been observed that there is no significant difference of the transcriptional level of flhF between normal protein secretion and high-level expression of the α-amylase AmyQ from Bacillus amyloliquefaciens. flhF and its neighboring genes, such as yxlH, are involved in cell motility. flhF has a minor role, as the single mutation in flhF does not affect cell motility whereas the combined flhF and ylxH mutations affect motility more severely than a single ylxH mutation. In addition, the function of FlhF can be replaced by FtsY but not vice versa [124]. The preliminary X-ray crystallographic analysis of FlhF fragments described its possible importance for the assembly of polar flagella but not in protein secretion [5].

Transferring and passing. Second, the translocation of proteins across the membrane is performed through the channel of the SecYEG translocase complex and the accessory protein SecDF. At the same time, SecA catalyzes the hydrolysis of ATP, providing energy for crossing the membrane.

The function of SecYEG in the translocation channel is thought to be similar to that of Escherichia coli [121]. The mechanism of SecA-dependent translocation of secretory protein through SecYEG channels was reported by Kudva et al. [56]. The overexpression of a constructed artificial secYEG operon resulting in an increased number of translocons enhances the secretion amounts of the active exoprotein α-amylase [75]. The plasmid pAX01 containing secY, secE, and secG, which are under control of the xyloseinducible PxylA promoter, integrates into the bacterial chromosome via a double crossing-over event at the lacA locus. SecDF is proposed to pull the pre-protein out of the SecYEG channel. Evidence resulted from the comparison of the structure but still lacks biochemical testing [68].

The homotetrameric SecB is employed as a targeting factor that binds to the core regions of the post-translated pre-proteins. It t argets t hem to t he S ecA subunit o f the translocase in order to maintain them in an unfolded translocation-competent state in gram-negative bacteria. Since Bacillus subtilis lacks the homolog to SecB, the 12 kDa homodimeric CsaA fulfills a similar function to SecB in Escherichia coli [96] for a certain subset of pre-proteins, including PhoB [76] and YvaY [62]. Defect of SecA in Escherichia coli, affecting normal growth and protein secretion, can be suppressed by chaperone-like activities of Bacillus subtilis’ CsaA. This is consistent with the fact that CsaA is significantly induced to combat secretion stress when α-amylase is highly expressed [41]. It remains to be an ongoing subject of study to elucidate the exact mechanism of CsaA, even after determination of crystallographic details.

SecA, a dual functional element, acts as a co-chaperone as well as the motor component of the Sec translocase. Even though the structure of SecA has been reviewed by Papanikou et al. [83] and Rocak and Linder [92] in detail, the mechanism of its function is still unclear. Here, we summarize some research on it. Zimmer et al. [131] reported on different dimers of SecA, indicating that they are all in equilibrium with monomers and do not differ dramatically in their stability. Moreover, and regarding the X-ray structure of Bacillus subtilis SecA, they assume that this monomer is required for protein translocation. One commonly known protein-conducting channel consisting of SecA, SecYEG, and SecDF has high affinity. The other low-affinity channel is composed by SecA alone. Preproteins can be transformed into the high-affinity channels owing to the lower specifity and less efficiency of SecA oligomers [111]. Both the tail-to-head tandem dimer SecAA and the tail-to-head SecAAA trimer complement a temperature-sensitive SecA mutant, even though these purified dimers and trimers show low-level intrinsic ATPase, similar to soluble monomeric SecA. Nevertheless, their activity can be improved to the same extent as the wild-type SecA protomer when stimulated by liposomes. It also was suggested that SecA may exist and function as a hexamer in the membrane.

In order to make full use of SecA and to improve secretion, many efforts have been made, either changing SecA itself or other components interacting with SecA. Kakeshita et al. [44] examined the influence of modifications in the C-terminal region of SecA on recombinant protein production in Bacillus subtilis. The C-terminal domain of SecA includes the poorly conserved C-terminal linker (CTL) and the highly conserved SecB-binding site, even though there is no SecB in Bacillus subtilis. The CTD of SecA seems not to be essential for heterologous protein secretion. When the 66 amino acids of SecA’s C-terminal are deleted, there is no difference between the SecA mutant and the wild-type strain. Not only is the cell viability normal, but the activity of the thermostable heterologous alkaline cellulase (Egl-237) also increases by 83%. Human interferon-α (hIFN-α2b) productivity increases about 120%. Furthermore, Diao et al. [20] constructed a co-expression system of SecB from Escherichia coli and a hybrid SecA in Bacillus subtilis. At the same time, the last 32 amino acids at the C-terminal of SecA were replaced by the corresponding part from Escherichia coli’s SecA, thus admitting the artificial SecA combined to SecB. A dramatic improvement of exported protein production took place when assessing the efficiency of the system. All these studies pave the way for future understanding of the mysterious translocase.

Removal and degradation. Third, the cleavage of Nterminal signal peptide sequences is accomplished by SPPases. Then, SppA and TepA are r ecruited for the subsequent degradation of the cleaved signal peptides. Otherwise, the cleaved signal peptides may become inhibitors to other protein translocation.

The SPPases in gram-positive species of Bacillus subtilis are divided into two types with respect to their distinct substrate specificity: type I (Lep) for most secretion proteins and type II (LspA) for lipoproteins. Five of seven sip genes for type I signal peptidase (denoted as sipS, sipT, sipU, sipV, and sipW) have been identified to locate on the chromosome. The presence of either SipS or SipT is sufficient for processing of precursor proteins and cell viability [104]. Multiple SPases may serve to guarantee a flexible capacity for protein secretion under various conditions. For example, SipW is a bifunctional signal peptidase, and is not only requested for the processing of TasA, which is a spore-associated protein [105] but also controls surface adherent biofilm formation [100]. The other two genes encoding SipP are part of homologous structural modules presented on at least two cryptic plasmids, pTA1015 and pTA1040. Their transcription product SipP can functionally replace SipS and SipT [106]. Since type II is responsible for lipoprotein cleavage, however, processing of lipoproteins by LspA is not strictly required for lipoprotein function [52]. The subject of SPase I and SPase II deletions has been reviewed by Tjalsm et al. [102] in order to further elucidate the effects of the signal peptidases.

The signal peptide peptidase SppA, a membrane-bound protein, is considered to be merely required under special conditions, such as hypersecretion [9]. Recent studies showed that SppA self-processes its own C-terminus, and proposed a model for how the C -terminus of SppA may function in the regulation of this membrane-bound selfcompartmentalized protease [78]. Unlike SppA, TepA seems to be located in the cytoplasmic matrix. TepA depletion has a significant effect on pre-protein secretion across the membrane and subsequent processing not only under conditions of hypersecretion [9].

Modification and maturing. Fourth, the translocational protein modification, folding process, and following procedures are catalyzed by factors including chaperones, pro-peptides, the peptidyl-prolyl cis/trans isomerase PrsA, disulfide isomerases BdbBCD, metal ions, and a quality control system.

Bacillus subtilis has intracellular and extracytoplasmic chaperones mediating protein folding, decreasing aggregation, and maintaining pre-proteins in translocation-competent conformations. The intracellular chaperone series of GroE and DnaK are regulated by the groE operon (groES–groEL) and the dnaK operon (hrcA-grpE-dnaK-dnaJ-yqeT-yqeU-yqeV) [36, 59]. Except for these functions and summarized in publications [90] and patents [79] (mentioned above), it was reported that the chaperones DnaK and GroEL are recruited to membranes for early membrane protection under short-term ethanol stress conditions [95]. In Bacillus subtilis, CsaA seems to act as a chaperone by preventing their aggregation, in contrast to SecB in Escherichia coli.

The membrane-anchored lipoprotein PrsA belongs to the extracellular chaperones and is indispensable for normal growth by facilitating the mature protein to fold into its stable and correct conformation. Overexpression of the substrate-specific chaperone PrsA results in a dramatic increase of α-amylase production [14, 109]. Consistent observations occur when AmyE pro-peptide is fused to human interferon-β. In this case, and under co-expression of PrsA, the recombinant protein production increases by more than 1.5-fold [45]. In addition to this, it is suggested that the folding of penicillin-binding protein is dependent on PrsA, as the deletion of PrsA leads to unstable PBP2a, PBP2b, PBP3, and PBP4 [40]. Reporter assays showed that extracytoplasmic chaperones act after intracellular chaperones; that is, in a sequential manner when assessing the MH-1 (fibrin) SCA production [117].

The formation of disulfide bonds, catalyzed by thioldisulfide oxidoreductases BdbBCD, is necessary for the stability and activity of many exo-proteins [10]. Overexpression of these thiol-disulfide oxidoreductases could improve secretion of proteins containing certain disulfide bonds, as reviewed by Nijland and Kuipers [79]. However, this increase was not significant, thus leading to the occurrence of a novel approach. To achieve an increased production, one should combine depletion of the major cytoplasmic reductase TrxA, introduction of the heterologous oxidase DsbA from Staphylococcus carnosus, and addition of redox-active compounds to the growth medium [53].

Partly misfolded proteins may aggregate into inclusion bodies both in and out of the cytoplasm. The amount of these cytoplasmic insoluble aggregates can be decreased by ClpCP, ClpXP, and ClpEP [37]. In Bacillus subtilis, the genes coding for ClpC, ClpE, and ClpP are members of the CtsR regulon (class of three stress gene repressors) except for ClpX [19]. ClpP is essential for intracellular protein quality control [50]. ClpE is involved in global protein disaggregation [71].

The quality control proteases consist of the two membraneanchored proteins HtrAB and the cell wall-bound protease WprA [17]. All of them are considered to be responsible for inducing degradation of incorrectly folded proteins that had passed the membrane. The two-component system CssRS (control secretion stress regulator and sensor) controls the expression of the htrA and htrB genes in response to the stress induction due to the overexpression of both homologous and heterologous secretory proteins [39]. However, recent observations suggest that HtrAB, being indeed the substrate of multiple extracellular proteases [55], is crucial for maintaining the integrity of the Bacillus subtilis cell even under non-stress conditions [54].

Tat Pathway

Different from the Sec pathway, the Tat pathway transports fully folded proteins containing a conserved region of twin-arginine in the signal peptide sequences. The Tat pathway facilitates proteins that have been folded too fast and tightly in the cytoplasm to be compatible to the Sec pathway. At the same time, both cofactors and complex redox enzymes, which are involved in the respiratory chain and insert into the cytosol prior to transport, are necessary.

The basic procedures include these steps: First and before translocation, pre-proteins fold in the cytoplasm with the help of cofactors. Second, owing to the presence of a pH gradient across the cytosolic membrane, pre-proteins are transported through the Tat translocase [91]. Third, after translocation, type I Spases cleave signal peptides. The remaining proteins secrete into the medium. To elucidate the mechanisms of this pathway, more attention needs to be paid to the details concerning processing, proofreading, and quality control of Tat substrates.

Unlike the large amounts of proteins secreted via Sec translocons, only three substrates of the Tat pathway in Bacillus subtilis have been identified: PhoD, YwbN, and QcrA. Bacillus subtilis contains two tatC genes, denoted tatCd and tatCy. Both tatCd and tatCy are preceded by tatA genes, denoted tatAd and tatAy, respectively. TatAdCd and TatAyCy act as the Tat machinery with the function of secreting the twin-arginine pre-proteins PhoD and YwbN, respectively [42]. The membrane protein QcrA is a novel Tat pathway-dependent substrate, unveiled by quantitative proteomic analysis [29].

The 3D model of TatAd of Bacillus subtilis observed from electron microscopy by Beck et al. [7] suggests that TatAd is not able to form a transporting channel by itself because of its small diameter of the TatAd pore. The solution NMR structure of TatAd showed a transmembrane (TMS) and an amphipathic helix (APH) forming an L-shaped structure and a largely unstructured C-terminal tail containing a densely charged region (DCR) [38, 110]. Walther et al. [110] proposed a charge zipper mechanism based on the complementarity between the DCR and APH; thus, the protein being “zippered up” by seven salt bridges forming a hairpin that crosses the lipid bilayer. During the transport process, TatA is an early interacting partner of functionally targeted Tat substrates [26]. TatC plays a role as insertase for twin-arginine signal peptides [25]. In contrast to Bacillus subtilis, gram-negative bacteria possess Tat translocases comprising the three membrane proteins TatA, TatB, and TatC. Previous and recent studies have showed that beyond recognizing twin-arginine signal peptides, TatC mediates the transmembrane insertion of a signal peptide. This is followed by the translocation of the cleavage sites across the bilayer with a concerted regulation of TatB [25].

The Tat pathway has become an increasingly popular transporting pathway in virtue of exporting fully folded proteins of interest into the cultural medium, decreasing the costs of purification. Furthermore, much attention has been paid to the investigations of biotechnology application. For instance, Yang et al. [122] expressed the methyl parathion hydrolase (MPH), an organophosphate (OP)- hydrolyzing enzyme, adding the twin-arginine signal peptide of trimethylamine N-oxide reductase (TorA) from Escherichia coli to MPH in Bacillus subtilis [51, 65]. The subsequent investigation exhibited an effective degradation of OPs on vegetables by MPHs, which are exported by the Tat pathway, exclusively. In addition, neither cell growth nor cell viability is inhibited by the secretion of MPH. It was reported by Xia et al. [119] that the cytoplasmic thermostable β-galactosidase from Geobacillus stearothermophilus carrying a twin-arginine signal peptide of Bacillus subtilis phosphodiesterase shows a relatively high exporting efficiency in Bacillus subtilis through the Tat pathway. This efficiency decreases when lacking the signal peptide sequence at the N-terminus. Apart from the useful expression system of Bacillus subtilis itself, the TatAdCd system plays a key role for substitutes of the TatABC translocases in Escherichia coli. After the signal peptide of TorA is placed ahead of green fluorescent protein (GFP), the amount of extracellular GFP increases owing to the leaking outer membrane. Nevertheless, cell viability is not affected [3]. These findings provide potent transport approach methods for recombinant heterologous proteins in large-scale production.

ABC Pathway

ATP-binding cassette (ABC) transporters comprise a comprehensive family of proteins existing in all living organisms, ranging from bacteria to humans. Despite that all ABC transporters can either import or export a wide variety of structurally and functionally unrelated substrates, they all share one characterization: They all include two transmembrane domains (TMDs), defining the substratebinding site, and two soluble nucleotide-binding domains (NBDs) acting as the motor domain. Unlike the diverse translocator units of the TMD, several conserved sequence motifs have been identified in the NBD [35]. However, the molecular mechanisms of ABC transporter proteins remain unclear.

These ABC transporters implement their functions in different aspects. Here, we list a part of recent advances in this field: Bacillus subtilis holds a set of five osmotically inducible osmoprotectant uptake systems, the Opu family of ABC transporters. Among them, OpuA is the main glycine betaine uptake system. The intracellular solute pool plays a key role in the osmotic control of OpuA expression [34]. The crystal structure of the OpuBC/choline complex, which is the substrate-binding protein of the OpuB transporters, provides a rational explanation for the observed choline specificity of the OpuB ABC importer in vivo [89]. Different from OpuB, structures of the substratebinding proteins give insight into the multi-compatible solute-binding specificities of OpuC [22].

There are also many response regulators and operonic kinase-regulator pairs in Bacillus subtilis. Close relationships exist between some operons encoding two-component systems and genes for ABC transporter homologs. The natAB operon, encoding a two-gene ABC transporter and involved in sodium ion extrusion, has been identified to be regulated by the NatKR two-component system [80]. The FtsEX ABC transporter regulates cellular differentiation in initiating sporulation [28]. YheI/YheH form a new heterodimeric multidrug ABC transporter and are predicted to be possibly responsible in multiple antibiotic resistance [107].

 

Transformation Methods and Genetic Engineering Strategies

Transformation Methods

All of the genetic engineering strategies are based on the development of high-efficiency transformation methods. The usual approaches comprise the competent transformation method [98], the protoplast transformation method [13], the protoplast electroporation method [57], the alkali metal ions method [4], and different kinds of electroporation methods and strategies based on the restriction modification system.

According to the competent transformation method, Bacillus subtilis enters a physiological state after incubation in GMI and GMII media, permitting the uptake of exogenous DNA. This process is controlled by a complex signal transduction cascade. This classical method is easy to manipulate, does not require special machines, and also results in high transformation frequencies.

The protoplast transformation method, nevertheless, is time consuming and labor intensive, owing to the protoplast preparation by degrading the cell wall and requiring experience on manipulation. In addition, cell regeneration takes several days. However, it is a useful transformation method, especially when the bacteria are difficult to be transformed owing to the physiological barrier.

Electroporation is a simple and widely used technique for transformation of various Bacillus species. Adding sorbitol, mannitol, or glycerol in hyperosmotic concentrations in the electroporation medium as well as in the growth and recovery media enables extremely high electric field strengths and improves transformation efficiencies [120]. Protection by high osmolarity media during electrobombardment improves cell survivability. In addition, trehalose [67], sorbitol and mannitol decrease damage of electric shocks to the cells when they are added to electroporation media. In Bacillus subtilis, this results in increased transformation efficiency by nearly 100-fold compared with conventional methods. Apart from adding these osmoticums in order to balance the high internal turgor pressure of bacterial cells, adding glycine and D/Lthreonine weakens cell walls even more. It is another strategy to elevate the electrotransformation efficiency, as well as cell-membrane fluidity when adding Tween 80 [128].

Moreover, the in-vivo-methylation strategy improves transformation efficiencies by methylation of exogenous plasmid DNA before transforming undomesticated strains or strains containing strict restriction and modification systems (RMS) [27]. Not only can the plasmids be methylated to deny the restriction enzymes, but also the restriction system in Bacillus can be modified. Deletion of two type I RMS of the hsdR loci creates a library of readily transformable [113] mutants of Bacillus licheniformis DSM13. This may also be applicable to Bacillus subtilis. Overexpression of the competence master regulator ComK is another alternative to make host strains become supercompetent cells. The transformation efficiency is highly improved when multimeric plasmids, generated by using prolonged overlap extension-PCR, are transformed [130]. There is also an approach based on conjugational plasmids, which is able to burden extra DNA and transmit it to another Bacillus [81].

Genetic Engineering Strategies

There has been much progress in the development of genetic modification strategies of Bacillus species (Dong and Zhang [21]). The operator-repressor system based on genetic engineering strategies includes CI repressors from Escherichia coli, bacteriophage λ and araR. araR is a negative regulator of the ara operon that can be induced by Larabinose in Bacillus subtilis as a counter-selection marker. upp and pyrF, encoding uracil phosphoribosyltransferase (UPRTase) and orotidine 5’-phosphate decarboxylase (OMPdecase), have been used as counter-selection markers. These genetic engineering strategies make use of the pyrimidine metabolism. In addition, there are genetic engineering strategies in different Bacillus species based on auxotrophy, site-specific recombination, toxic genes, thermosensitive plasmids, and transconjugation.

In conclusion, taken together, increasing information about Bacillus subtilis in both basic research and application demonstrates that this potential industrial bacterium will become a super-secreting cell factory through continuous research and understanding [47], particularly in terms of the Sec and Tat pathways. Further studies are still needed in order to understand the function of each element of the Sec secretory machinery and their interaction. Further focus on the Tat pathway seems indispensable, since it enables folded proteins to travel through the membrane, ensuring the bioactivity of products. Attracting further attention, the network between the different pathways allows robust bacteria to overcome unfavorable environmental conditions and pass the bottlenecks of secretion.

Apart from the advances in elucidation of complicated mechanisms and control of the Sec and Tat pathways, there may be other aspects to increase secretion efficiency (Fig. 3): the development of protease-deficiency strains, minimizing the unnecessary genome sequences [69, 70, 74], and optimization of metabolism engineering techniques. Harness of the systems and synthetic biological approaches can improve industrial production, with the combination of genomics and proteomics. In conclusion, it is crucial to investigate the mechanisms of the secretion pathways and the influencing factors in order to broaden the range of efficient industrial application of Bacillus subtilis.

Fig. 3.Scheme of strategies to improve the protein secretion production by Bacillus subtilis. RBS, ribosome-binding site; SP, signal peptide.

References

  1. Aiyar SE, Gourse RL, Ross W. 1998. Upstream A-tracts increase bacterial promoter activity through interactions with the RNA polymerase alpha subunit. Proc. Natl. Acad. Sci. USA 95: 14652-14657. https://doi.org/10.1073/pnas.95.25.14652
  2. Akopian D, Shen K , Zhang X, Shan SO. 2013. Signal recognition particle: an essential protein-targeting machine. Annu. Rev. Biochem. 82: 693-721. https://doi.org/10.1146/annurev-biochem-072711-164732
  3. Albiniak AM, Matos CF, Branston SD, Freedman RB, Keshavarz-Moore E, Robinson C. 2013. High-level secretion of a recombinant protein to the culture medium with a Bacillus subtilis twin-arginine translocation system in Escherichia coli. FEBS J. 280: 3810-3821. https://doi.org/10.1111/febs.12376
  4. Ano T, Kobayash iA, Shoda M. 1990. Transformation of Bacillus subtilis with the treatment by alkali cations. Biotechnol. Lett. 12: 99-104. https://doi.org/10.1007/BF01022423
  5. Bange G, Petzold G, Wild K, Sinning I. 2007. Expression, purification and preliminary crystallographic characterization of FlhF from Bacillus subtilis. Acta Crystallogr. Sect. F Struct. Biol. Cryst. Commun. 63: 449-451. https://doi.org/10.1107/S1744309107020180
  6. Bechhofer DH. 2009. Messenger RNA decay and maturation in Bacillus subtilis. Prog. Mol. Biol. Transl. Sci. 85: 231-273. https://doi.org/10.1016/S0079-6603(08)00806-4
  7. Beck D, Vasisht N, Baglieri J, Monteferrante CG, van Dijl JM, Robinson C, Smith CJ. 2013. Ultrastructural characterisation of Bacillus subtilis TatA complexes suggests they are too small to form homooligomeric translocation pores. Biochim. Biophys. Acta 1833: 1811-1819. https://doi.org/10.1016/j.bbamcr.2013.03.028
  8. Bhavsar AP, Zhao X, Brown ED. 2001. Development and characterization of a xylose-dependent system for expression of cloned genes in Bacillus subtilis: conditional complementation of a teichoic acid mutant. Appl. Environ. Microbiol. 67: 403-410. https://doi.org/10.1128/AEM.67.1.403-410.2001
  9. Bolhuis A, Matzen A, Hyyrylainen HL, Kontinen VP, Meima R, Chapuis J, et al. 1999. Signal peptide peptidaseand ClpP-like proteins of Bacillus subtilis required for efficient translocation and processing of secretory proteins. J. Biol. Chem. 274: 24585-24592. https://doi.org/10.1074/jbc.274.35.24585
  10. Bolhuis A, Venema G, Quax WJ, Bron S, van Dijl JM. 1999. Functional analysis of paralogous thiol-disulfide oxidoreductases in Bacillus subtilis. J. Biol. Chem. 274: 24531-24538. https://doi.org/10.1074/jbc.274.35.24531
  11. Brockmeier U, Caspers M, Freudl R, Jockwer A, Noll T, Eggert T. 2006. Systematic screening of all signal peptides from Bacillus subtilis: a powerful strategy in optimizing heterologous protein secretion in gram-positive bacteria. J. Mol. Biol. 362: 393-402. https://doi.org/10.1016/j.jmb.2006.07.034
  12. Caspers M, Brockmeier U, Degering C, Eggert T, Freudl R. 2010. Improvement of Sec-dependent secretion of a heterologous model protein in Bacillus subtilis by saturation mutagenesis of the N-domain of the AmyE signal peptide. Appl. Microbiol. Biotechnol. 86: 1877-1885. https://doi.org/10.1007/s00253-009-2405-x
  13. Chang S, Cohen SN. 1979. High frequency transformation of Bacillus subtilis protoplasts by plasmid DNA. Mol. Gen. Genet. 168: 111-115. https://doi.org/10.1007/BF00267940
  14. Chen J, Gai Y, Fu G, Zhou W, Zhang D, Wen J. 2014. Enhanced extracellular production of alpha-amylase in Bacillus subtilis by optimization of regulatory elements and over-expression of PrsA lipoprotein. Biotechnol. Lett. 37: 899-906. https://doi.org/10.1007/s10529-014-1755-3
  15. Chen PT, Shaw JF, Chao YP, Ho TH, Yu SM. 2010. Construction of chromosomally located T7 expression system for production of heterologous secreted proteins in Bacillus subtilis. J. Agric. Food Chem. 58: 5392-5399. https://doi.org/10.1021/jf100445a
  16. Chung YS, Breidt F, Dubnau D. 1998. Cell surface localization and processing of the ComG proteins, required for DNA binding during transformation of Bacillus subtilis. Mol. Microbiol. 29: 905-913. https://doi.org/10.1046/j.1365-2958.1998.00989.x
  17. Darmon E, Noone D, Masson A, Bron S, Kuipers OP, Devine KM, van Dijl JM. 2002. A novel class of heat and secretion stress-responsive genes is controlled by the autoregulated CssRS two-component system of Bacillus subtilis. J. Bacteriol. 184: 5661-5671. https://doi.org/10.1128/JB.184.20.5661-5671.2002
  18. Degering C, Eggert T, Puls M, Bongaerts J, Evers S, Maurer KH, Jaeger KE. 2010. Optimization of protease secretion in Bacillus subtilis and Bacillus licheniformis by screening of homologous and heterologous signal peptides. Appl. Environ. Microbiol. 76: 6370-6376. https://doi.org/10.1128/AEM.01146-10
  19. Derre I, Rapoport G, Msadek T. 1999. CtsR, a novel regulator of stress and heat shock response, controls clp and molecular chaperone gene expression in gram-positive bacteria. Mol. Microbiol. 31: 117-131. https://doi.org/10.1046/j.1365-2958.1999.01152.x
  20. Diao L, Dong Q, Xu Z, Yang S, Zhou J, Freudl R. 2012. Functional implementation of the posttranslational SecBSecA protein-targeting pathway in Bacillus subtilis. Appl. Environ. Microbiol. 78: 651-659. https://doi.org/10.1128/AEM.07209-11
  21. Dong H, Zhang D. 2014. Current development in genetic engineering strategies of Bacillus species. Microb. Cell Fact. 13: 63. https://doi.org/10.1186/1475-2859-13-63
  22. Du Y, Shi WW, He YX, Yang YH, Zhou CZ, Chen Y. 2011. Structures of the substrate-binding protein provide insights into the multiple compatible solute binding specificities of the Bacillus subtilis ABC transporter OpuC. Biochem. J. 436: 283-289. https://doi.org/10.1042/BJ20102097
  23. Ferreira LC, Ferreira RC, Schumann W. 2005. Bacillus subtilis as a tool for vaccine development: from antigen factories to delivery vectors. An. Acad. Bras. Cienc. 77: 113-124. https://doi.org/10.1590/S0001-37652005000100009
  24. Fischer KE, Bremer E. 2012. Activity of the osmotically regulated yqiHIK promoter from Bacillus subtilis is controlled at a distance. J. Bacteriol. 194: 5197-5208. https://doi.org/10.1128/JB.01041-12
  25. Frobel J, Rose P, Lausberg F, Blummel AS, Freudl R, Muller M. 2012. Transmembrane insertion of twin-arginine signal peptides is driven by TatC and regulated by TatB. Nat. Commun. 3: 1311. https://doi.org/10.1038/ncomms2308
  26. Frobel J, Rose P, Muller M. 2011. Early contacts between substrate proteins and TatA translocase component in twin-arginine translocation. J. Biol. Chem. 286: 43679-43689. https://doi.org/10.1074/jbc.M111.292565
  27. Gao C, Xue Y, Ma Y. 2011. Protoplast transformation of recalcitrant alkaliphilic Bacillus sp. with methylated plasmid DNA and a developed hard agar regeneration medium. PLoS One 6: e28148. https://doi.org/10.1371/journal.pone.0028148
  28. Garti-Levi S, Hazan R, Kain J, Fujita M, Ben-Yehuda S. 2008. The FtsEX ABC transporter directs cellular differentiation in Bacillus subtilis. Mol. Microbiol. 69: 1018-1028. https://doi.org/10.1111/j.1365-2958.2008.06340.x
  29. Goosens VJ, Otto A, Glasner C, Monteferrante CC, van der Ploeg R, Hecker M, et al. 2013. Novel twin-arginine translocation pathway-dependent phenotypes of Bacillus subtilis unveiled by quantitative proteomics. J. Proteome Res. 12: 796-807. https://doi.org/10.1021/pr300866f
  30. Haldenwang WG. 1995. The sigma factors of Bacillus subtilis. Microbiol. Rev. 59: 1-30.
  31. Hambraeus G, Karhumaa K, Rutberg B. 2002. A 5’ stemloop and ribosome binding but not translation are important for the stability of Bacillus subtilis aprE leader mRNA. Microbiology 148: 1795-1803. https://doi.org/10.1099/00221287-148-6-1795
  32. Helmann JD. 1995. Compilation and analysis of Bacillus subtilis sigma(a)-dependent promoter sequences - evidence for extended contact between RNA-polymerase and upstream promoter DNA. Nucleic Acids Res. 23: 2351-2360. https://doi.org/10.1093/nar/23.13.2351
  33. Heravi KM, Wenzel M, Altenbuchner J. 2011. Regulation of mtl operon promoter of Bacillus subtilis: requirements of its use in expression vectors. Microb. Cell Fact. 10: 83. https://doi.org/10.1186/1475-2859-10-83
  34. Hoffmann T, Wensing A, Brosius M, Steil L, Volker U, Bremer E. 2013. Osmotic control of opuA expression in Bacillus subtilis and its modulation in response to intracellular glycine betaine and proline pools. J. Bacteriol. 195: 510-522. https://doi.org/10.1128/JB.01505-12
  35. Hollenstein K, Dawson RJ, Locher KP. 2007. Structure and mechanism of ABC transporter proteins. Curr. Opin. Struct. Biol. 17: 412-418. https://doi.org/10.1016/j.sbi.2007.07.003
  36. Homuth G, Masuda S, Mogk A, Kobayashi Y, Schumann W. 1997. The dnaK operon of Bacillus subtilis is heptacistronic. J. Bacteriol. 179: 1153-1164. https://doi.org/10.1128/jb.179.4.1153-1164.1997
  37. Hoskins JR, Sharma S, Sathyanarayana BK, Wickner S. 2001. Clp ATPases and their role in protein unfolding and degradation. Adv. Protein Chem. 59: 413-429. https://doi.org/10.1016/S0065-3233(01)59013-0
  38. Hu Y, Zhao E, Li H, Xia B, Jin C. 2010. Solution NMR structure of the TatA component of the twin-arginine protein transport system from gram-positive bacterium Bacillus subtilis. J. Am. Chem. Soc. 132: 15942-15944. https://doi.org/10.1021/ja1053785
  39. Hyyrylainen HL, Bolhuis A, Darmon E, Muukkonen L, Koski P, Vitikainen M, et al. 2001. A novel two-component regulatory system in Bacillus subtilis for the survival of severe secretion stress. Mol. Microbiol. 41: 1159-1172. https://doi.org/10.1046/j.1365-2958.2001.02576.x
  40. Hyyrylainen HL, Marciniak BC, Dahncke K, Pietiainen M, Courtin P, Vitikainen M, et al. 2010. Penicillin-binding protein folding is dependent on the PrsA peptidyl-prolyl cis-trans isomerase in Bacillus subtilis. Mol. Microbiol. 77: 108-127. https://doi.org/10.1111/j.1365-2958.2010.07188.x
  41. Hyyrylainen HL, Sarvas M, Kontinen VP. 2005. Transcriptome analysis of the secretion stress response of Bacillus subtilis. Appl. Microbiol. Biotechnol. 67: 389-396. https://doi.org/10.1007/s00253-005-1898-1
  42. Jongbloed JD, Grieger U, Antelmann H, Hecker M, Nijland R, Bron S, van Dijl JM. 2004. Two minimal Tat translocases in Bacillus. Mol. Microbiol. 54: 1319-1325. https://doi.org/10.1111/j.1365-2958.2004.04341.x
  43. Jurgen B, Schweder T, Hecker M. 1998. The stability of mRNA from the gsiB gene of Bacillus subtilis is dependent on the presence of a strong ribosome binding site. Mol. Gen. Genet. 258: 538-545. https://doi.org/10.1007/s004380050765
  44. Kakeshita H, Kageyama Y, Ara K, Ozaki K, Nakamura K. 2010. Enhanced extracellular production of heterologous proteins in Bacillus subtilis b y deleting t h e C-terminal region of the SecA secretory machinery. Mol. Biotechnol. 46: 250-257. https://doi.org/10.1007/s12033-010-9295-0
  45. Kakeshita H, Kageyama Y, Endo K, Tohata M, Ara K, Ozaki K, Nakamura K. 2011. Secretion of biologicallyactive human interferon-beta by Bacillus subtilis. Biotechnol. Lett. 33: 1847-1852. https://doi.org/10.1007/s10529-011-0636-2
  46. Kang HK, Jang JH, Shim JH, Park JT, Kim YW, Park KH. 2010. Efficient constitutive expression of thermostable 4- alpha-glucanotransferase in Bacillus subtilis using dual promoters. World J. Microbiol. Biotechnol. 26: 1915-1918. https://doi.org/10.1007/s11274-010-0351-5
  47. Kang Z, Yang S, Du G, Chen J. 2014. Molecular engineering of secretory machinery components for high-level secretion of proteins in Bacillus species. J. Ind. Microbiol. Biotechnol. 41: 1599-1607. https://doi.org/10.1007/s10295-014-1506-4
  48. Kempf G, Wild K, Sinning I. 2014. Structure of the complete bacterial SRP Alu domain. Nucleic Acids Res. 42: 12284-12294. https://doi.org/10.1093/nar/gku883
  49. Kim JH, Lee BR, Lee YP. 2011. Secretory overproduction of the aminopeptidase from Bacillus licheniformis by a novel hybrid promoter in Bacillus subtilis. World J. Microbiol. Biotechnol. 27: 2747-2751. https://doi.org/10.1007/s11274-011-0749-8
  50. Kock H, Gerth U, Hecker M. 2004. The ClpP peptidase is the major determinant of bulk protein turnover in Bacillus subtilis. J. Bacteriol. 186: 5856-5864. https://doi.org/10.1128/JB.186.17.5856-5864.2004
  51. Kolkman MA, van der Ploeg R, Bertels M, van Dijk M, van der Laan J, van Dijl JM, Ferrari E. 2008. The twin-arginine signal peptide of Bacillus subtilis YwbN can direct either Tat- or Sec-dependent secretion of different cargo proteins: secretion of active subtilisin via the B. subtilis Tat pathway. Appl. Environ. Microbiol. 74: 7507-7513. https://doi.org/10.1128/AEM.01401-08
  52. Kontinen VP, Sarvas M. 1993. The PrsA lipoprotein is essential for protein secretion in Bacillus subtilis and sets a limit for high-level secretion. Mol. Microbiol. 8: 727-737. https://doi.org/10.1111/j.1365-2958.1993.tb01616.x
  53. Kouwen TR, Dubois JY, Freudl R, Quax WJ, van Dijl JM. 2008. Modulation of thiol-disulfide oxidoreductases for increased production of disulfide-bond-containing proteins in Bacillus subtilis. Appl. Environ. Microbiol. 74: 7536-7545. https://doi.org/10.1128/AEM.00894-08
  54. Krishnappa L, Dreisbach A, Otto A, Goosens VJ, Cranenburgh RM, Harwood CR, et al. 2013. Extracytoplasmic proteases determining the cleavage and release of secreted proteins, lipoproteins, and membrane proteins in Bacillus subtilis. J. Proteome Res. 12: 4101-4110. https://doi.org/10.1021/pr400433h
  55. Krishnappa L, Monteferrante CG, Neef J, Dreisbach A, van Dijl JM. 2014. Degradation of extracytoplasmic catalysts for protein folding in Bacillus subtilis. Appl. Environ. Microbiol. 80: 1463-1468. https://doi.org/10.1128/AEM.02799-13
  56. Kudva R, Denks K, Kuhn P, Vogt A, Muller M, Koch HG. 2013. Protein translocation across the inner membrane of gram-negative bacteria: the Sec and Tat dependent protein transport pathways. Res. Microbiol. 164: 505-534. https://doi.org/10.1016/j.resmic.2013.03.016
  57. Kusaoke H, Hayashi Y, Kadowaki Y, Kimoto H. 1989. Optimum conditions for electric pulse-mediated gene-transfer to Bacillus subtilis cells. Agric. Biol. Chem. 53: 2441-2446. https://doi.org/10.1271/bbb1961.53.2441
  58. Lee SJ, Pan JG, Park SH, Choi SK. 2010. Development of a stationary phase-specific autoinducible expression system in Bacillus subtilis. J. Biotechnol. 149: 16-20. https://doi.org/10.1016/j.jbiotec.2010.06.021
  59. Li M, Wong SL. 1992. Cloning and characterization of the groESL operon from Bacillus subtilis. J. Bacteriol. 174: 3981-3992. https://doi.org/10.1128/jb.174.12.3981-3992.1992
  60. Li W, Li HX, Ji SY, Li S, Gong YS, Yang MM, Chen YL. 2007. Characterization of two temperature-inducible promoters newly isolated from B. subtilis. Biochem. Biophys Res. Commun. 358: 1148-1153. https://doi.org/10.1016/j.bbrc.2007.05.064
  61. Li W, Zhou X, Lu P. 2004. Bottlenecks in the expression and secretion of heterologous proteins in Bacillus subtilis. Res. Microbiol. 155: 605-610. https://doi.org/10.1016/j.resmic.2004.05.002
  62. Linde D, Volkmer-Engert R, Schreiber S, Muller JP. 2003. Interaction of the Bacillus subtilis chaperone CsaA with the secretory protein YvaY. FEMS Microbiol. Lett. 226: 93-100. https://doi.org/10.1016/S0378-1097(03)00578-0
  63. Ling LF, Zi RX, Wei FL, Jiang BS, Ping L, Chun XH. 2007. Protein secretion pathways in Bacillus subtilis: implication for optimization of heterologous protein secretion. Biotechnol. Adv. 25: 1-12. https://doi.org/10.1016/j.biotechadv.2006.08.002
  64. Liu L, Liu Y, Shin HD, Chen RR, Wang NS, Li J, et al. 2013. Developing Bacillus spp. as a cell factory for production of microbial enzymes and industrially important biochemicals in the context of systems and synthetic biology. Appl. Microbiol. Biotechnol. 97: 6113-6127. https://doi.org/10.1007/s00253-013-4960-4
  65. Liu R, Zuo Z, Xu Y, Song C, Jiang H, Qiao C, et al. 2014. Twin-arginine signal peptide of Bacillus subtilis YwbN can direct Tat-dependent secretion of methyl parathion hydrolase. J. Agric. Food Chem. 62: 2913-2918. https://doi.org/10.1021/jf405694n
  66. Liu SL, Du K. 2012. Enhanced expression of an endoglucanase in Bacillus subtilis by using the sucroseinducible sacB promoter and improved properties of the recombinant enzyme. Protein Expr. Purif. 83: 164-168. https://doi.org/10.1016/j.pep.2012.03.015
  67. Lu YP, Zhang C, Lv FX, Bie XM, Lu ZX. 2012. Study on the electro-transformation conditions of improving transformation efficiency for Bacillus subtilis. Lett. Appl. Microbiol. 55: 9-14. https://doi.org/10.1111/j.1472-765X.2012.03249.x
  68. Lycklama ANJA, Driessen AJ. 2012. The bacterial Sectranslocase: structure and mechanism. Philos. Trans. R. Soc. Lond. B Biol. Sci. 367: 1016-1028. https://doi.org/10.1098/rstb.2011.0201
  69. Manabe K, Kageyama Y, Morimoto T, Ozawa T, Sawada K, Endo K, et al. 2011. Combined effect of improved cell yield and increased specific productivity enhances recombinant enzyme production in genome-reduced Bacillus subtilis strain MGB874. Appl. Environ. Microbiol. 77: 8370-8381. https://doi.org/10.1128/AEM.06136-11
  70. Manabe K, Kageyama Y, Morimoto T, Shimizu E, Takahashi H, Kanaya S, et al. 2013. Improved production of secreted heterologous enzyme in Bacillus subtilis strain MGB874 via modification of glutamate metabolism and growth conditions. Microb. Cell Fact. 12: 18. https://doi.org/10.1186/1475-2859-12-18
  71. Miethke M, Hecker M, Gerth U. 2006. Involvement of Bacillus subtilis ClpE in CtsR degradation and protein quality control. J. Bacteriol. 188: 4610-4619. https://doi.org/10.1128/JB.00287-06
  72. Ming YM, Wei ZW, Lin CY, Sheng GY. 2010. Development of a Bacillus subtilis expression system using the improved Pglv promoter. Microb. Cell Fact. 9: 55. https://doi.org/10.1186/1475-2859-9-55
  73. Moran CP Jr, Johnson WC, Losick R. 1982. Close contacts between sigma 37-RNA polymerase and a Bacillus subtilis chromosomal promoter. J. Mol. Biol. 162: 709-713. https://doi.org/10.1016/0022-2836(82)90399-0
  74. Morimoto T, Kadoya R, Endo K, Tohata M, Sawada K, Liu S, et al. 2008. Enhanced recombinant protein productivity by genome reduction in Bacillus subtilis. DNA Res. 15: 73-81. https://doi.org/10.1093/dnares/dsn002
  75. Mulder KC, Bandola J, Schumann W. 2013. Construction of an artificial secYEG operon allowing high level secretion of alpha-amylase. Protein Expr. Purif. 89: 92-96. https://doi.org/10.1016/j.pep.2013.02.008
  76. Muller JP, Ozegowski J, Vettermann S, Swaving J, Van Wely KH, Driessen AJ. 2000. Interaction of Bacillus subtilis CsaA with SecA and precursor proteins. Biochem. J. 348: 367-373. https://doi.org/10.1042/bj3480367
  77. Na D, Lee D. 2010. RBSDesigner: software for designing synthetic ribosome binding sites that yields a desired level of protein expression. Bioinformatics 26: 2633-2634. https://doi.org/10.1093/bioinformatics/btq458
  78. Nam SE, Paetzel M. 2013. Structure of signal peptide peptidase A with C-termini bound in the active sites: insights into specificity, self-processing, and regulation. Biochemistry 52: 8811-8822. https://doi.org/10.1021/bi4011489
  79. Nijland R, Kuipers OP. 2008. Optimization of protein secretion by Bacillus subtilis. Recent Pat. Biotechnol. 2: 79-87. https://doi.org/10.2174/187220808784619694
  80. Ogura M, Tsukahara K, Hayashi K, Tanaka T. 2007. The Bacillus subtilis NatK-NatR two-component system regulates expression of the natAB operon encoding an ABC transporter for sodium ion extrusion. Microbiology 153: 667-675. https://doi.org/10.1099/mic.0.2006/003673-0
  81. Ohtani N, Sato M, Tomita M, Itaya M. 2008. Restriction on conjugational transfer of pLS20 in Bacillus subtilis 168. Biosci. Biotechnol. Biochem. 72: 2472-2475. https://doi.org/10.1271/bbb.80315
  82. Paccez JD, Luiz WB, Sbrogio-Almeida ME, Ferreira RC, Schumann W, Ferreira LC. 2006. Stable episomal expression system under control of a stress inducible promoter enhances the immunogenicity of Bacillus subtilis as a vector for antigen delivery. Vaccine 24: 2935-2943. https://doi.org/10.1016/j.vaccine.2005.12.013
  83. Papanikou E, Karamanou S, Baud C, Frank M, Sianidis G, Keramisanou D, et al. 2005. Identification of the preprotein binding domain of SecA. J. Biol. Chem. 280: 43209-43217. https://doi.org/10.1074/jbc.M509990200
  84. Peterson JH, Woolhead CA, Bernstein HD. 2003. Basic amino acids in a distinct subset of signal peptides promote interaction with the signal recognition particle. J. Biol. Chem. 278: 46155-46162. https://doi.org/10.1074/jbc.M309082200
  85. Phan TT, Nguyen HD, Schumann W. 2010. Establishment of a simple and rapid method to screen for strong promoters in Bacillus subtilis. Protein Expr. Purif. 71: 174-178. https://doi.org/10.1016/j.pep.2009.11.010
  86. Phan TT, Nguyen HD, Schumann W. 2012. Development of a strong intracellular expression system for Bacillus subtilis by optimizing promoter elements. J. Biotechnol. 157: 167-172. https://doi.org/10.1016/j.jbiotec.2011.10.006
  87. Phan TT, Nguyen HD, Schumann W. 2013. Construction of a 5’-controllable stabilizing element (CoSE) for over-production of heterologous proteins at high levels in Bacillus subtilis. J. Biotechnol. 168: 32-39. https://doi.org/10.1016/j.jbiotec.2013.07.031
  88. Phan TT, Schumann W. 2007. Development of a glycineinducible expression system for Bacillus subtilis. J. Biotechnol. 128: 486-499. https://doi.org/10.1016/j.jbiotec.2006.12.007
  89. Pittelkow M, Tschapek B, Smits SH, Schmitt L, Bremer E. 2011. The crystal structure of the substrate-binding protein OpuBC from Bacillus subtilis in complex with choline. J. Mol. Biol. 411: 53-67. https://doi.org/10.1016/j.jmb.2011.05.037
  90. Pohl S, Harwood CR. 2010. Heterologous protein secretion by Bacillus species from the cradle to the grave. Adv. Appl. Microbiol. 73: 1-25. https://doi.org/10.1016/S0065-2164(10)73001-X
  91. Pop O, Martin U, Abel C, Muller JP. 2002. The twinarginine signal peptide of PhoD and the TatAd/Cd proteins of Bacillus subtilis form an autonomous Tat translocation system. J. Biol. Chem. 277: 3268-3273. https://doi.org/10.1074/jbc.M110829200
  92. Rocak S, Linder P. 2004. DEAD-box proteins: the driving forces behind RNA metabolism. Nat. Rev. Mol. Cell Biol. 5: 232-241. https://doi.org/10.1038/nrm1335
  93. Salis HM, Mirsky EA, Voigt CA. 2009. Automated design of synthetic ribosome binding sites to control protein expression. Nat. Biotechnol. 27: 946-950. https://doi.org/10.1038/nbt.1568
  94. 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.
  95. Seydlova G, Halada P, Fiser R, Toman O, Ulrych A, Svobodova J. 2012. DnaK and GroEL chaperones are recruited to the Bacillus subtilis membrane after short-term ethanol stress. J. Appl. Microbiol. 112: 765-774. https://doi.org/10.1111/j.1365-2672.2012.05238.x
  96. Shapova YA, Paetzel M. 2007. Crystallographic analysis of Bacillus subtilis CsaA. Acta Crystallogr. D Biol. Crystallogr. 63: 478-485. https://doi.org/10.1107/S0907444907005045
  97. Sharp JS, Bechhofer DH. 2003. Effect of translational signals on mRNA decay in Bacillus subtilis. J. Bacteriol. 185: 5372-5379. https://doi.org/10.1128/JB.185.18.5372-5379.2003
  98. Spizizen J. 1958. Transformation of biochemically deficient strains of Bacillus subtilis by deoxyribonucleate. Proc. Natl. Acad. Sci. USA 44: 1072-1078. https://doi.org/10.1073/pnas.44.10.1072
  99. Terpe K. 2006. Overview of bacterial expression systems for heterologous protein production: from molecular and biochemical fundamentals to commercial systems. Appl. Microbiol. Biotechnol. 72: 211-222. https://doi.org/10.1007/s00253-006-0465-8
  100. Terra R, Stanley-Wall NR, Cao G, Lazazzera BA. 2012. Identification of Bacillus subtilis SipW as a bifunctional signal peptidase that controls surface-adhered biofilm formation. J. Bacteriol. 194: 2781-2790. https://doi.org/10.1128/JB.06780-11
  101. Thuy Le AT, Schumann W. 2007. A novel cold-inducible expression system for Bacillus subtilis. Protein Expr. Purif. 53: 264-269. https://doi.org/10.1016/j.pep.2006.12.023
  102. Tjalsma H, Antelmann H, Jongbloed JD, Braun PG, Darmon E, Dorenbos R, et al. 2004. Proteomics of protein secretion by Bacillus subtilis: separating the “secrets” of the secretome. Microbiol. Mol. Biol. Rev. 68: 207-233. https://doi.org/10.1128/MMBR.68.2.207-233.2004
  103. Tjalsma H, Bolhuis A, Jongbloed JD, Bron S, van Dijl JM. 2000. Signal peptide-dependent protein transport in Bacillus subtilis: a genome-based survey of the secretome. Microbiol. Mol. Biol. Rev. 64: 515-547. https://doi.org/10.1128/MMBR.64.3.515-547.2000
  104. Tjalsma H, Bolhuis A, van Roosmalen ML, Wiegert T, Schumann W, Broekhuizen CP, et al. 1998. Functional analysis of the secretory precursor processing machinery of Bacillus subtilis: identification of a eubacterial homolog of archaeal and eukaryotic signal peptidases. Genes Dev. 12: 2318-2331. https://doi.org/10.1101/gad.12.15.2318
  105. Tjalsma H, Stover AG, Driks A, Venema G, Bron S, van Dijl JM. 2000. Conserved serine and histidine residues are critical for activity of the ER-type signal peptidase SipW of Bacillus subtilis. J. Biol. Chem. 275: 25102-25108. https://doi.org/10.1074/jbc.M002676200
  106. Tjalsma H, van den Dolder J, Meijer WJ, Venema G, Bron S, van Dijl JM. 1999. The plasmid-encoded signal peptidase SipP can functionally replace the major signal peptidases SipS and SipT of Bacillus subtilis. J. Bacteriol. 181: 2448-2454.
  107. Torres C, Galian C, Freiberg C, Fantino JR, Jault JM. 2009. The YheI/YheH heterodimer from Bacillus subtilis is a multidrug ABC transporter. Biochim. Biophys. Acta 1788: 615-622. https://doi.org/10.1016/j.bbamem.2008.12.012
  108. Vellanoweth RL, Rabinowitz JC. 1992. The influence of ribosome-binding-site elements on translational efficiency in Bacillus subtilis and Escherichia coli in vivo. Mol. Microbiol. 6: 1105-1114. https://doi.org/10.1111/j.1365-2958.1992.tb01548.x
  109. Vitikainen M, Lappalainen I, Seppala R, Antelmann H, Boer H, Taira S, et al. 2004. Structure-function analysis of PrsA reveals roles for the parvulin-like and flanking Nand C-terminal domains in protein folding and secretion in Bacillus subtilis.J. Biol. Chem. 279: 19302-19314. https://doi.org/10.1074/jbc.M400861200
  110. Walther TH, Gottselig C, Grage SL, Wolf M, Vargiu AV, Klein MJ, et al. 2013. Folding and self-assembly of the TatA translocation pore based on a charge zipper mechanism. Cell 152: 316-326. https://doi.org/10.1016/j.cell.2012.12.017
  111. Wang H, Ma Y, Hsieh YH, Yang H, Li M, Wang B, Tai PC. 2014. SecAAA trimer is fully functional as SecAA dimer in the membrane: existence of higher oligomers? Biochem. Biophys. Res. Commun. 447: 250-254. https://doi.org/10.1016/j.bbrc.2014.03.116
  112. Wang J, Ai X, Mei H, Fu Y, Chen B, Yu Z, He J. 2013. High-throughput identification of promoters and screening of highly active promoter-5’-UTR DNA region with different characteristics from Bacillus thuringiensis. PLoS One 8: e62960. https://doi.org/10.1371/journal.pone.0062960
  113. Waschkau B, Waldeck J, Wieland S, Eichstadt R, Meinhardt F. 2008. Generation of readily transformable Bacillus licheniformis mutants. Appl. Microbiol. Biotechnol. 78: 181-188. https://doi.org/10.1007/s00253-007-1278-0
  114. Wenzel M, Muller A, Siemann-Herzberg M, Altenbuchner J. 2011. Self-inducible Bacillus subtilis expression system for reliable and inexpensive protein production by high-celldensity fermentation. Appl. Environ. Microbiol. 77: 6419-6425. https://doi.org/10.1128/AEM.05219-11
  115. Westers L, Westers H, Quax WJ. 2004. Bacillus subtilis as cell factory for pharmaceutical proteins: a biotechnological approach to optimize the host organism. Biochim. Biophys. Acta 1694: 299-310. https://doi.org/10.1016/j.bbamcr.2004.02.011
  116. Wild K, Halic M, Sinning I, Beckmann R. 2004. SRP meets the ribosome. Nat. Struct. Mol. Biol. 11: 1049-1053. https://doi.org/10.1038/nsmb853
  117. Wu SC, Yeung JC, Duan Y, Ye R, Szarka SJ, Habibi HR, Wong SL. 2002. Functional production and characterization of a fibrin-specific single-chain antibody fragment from Bacillus subtilis: effects of molecular chaperones and a wallbound protease on antibody fragment production. Appl. Environ. Microbiol. 68: 3261-3269. https://doi.org/10.1128/AEM.68.7.3261-3269.2002
  118. Wu SM, Feng C, Zhong J, Huan LD. 2011. Enhanced production of recombinant nattokinase in Bacillus subtilis by promoter optimization. World J. Microbiol. Biotechnol. 27: 99-106. https://doi.org/10.1007/s11274-010-0432-5
  119. Xia Y, Zhao J, Chen H, Liu X, Wang Y, Tian F, et al. 2010. Extracellular secretion in Bacillus subtilis of a cytoplasmic thermostable beta-galactosidase from Geobacillus stearothermophilus. J. Dairy Sci. 93: 2838-2845. https://doi.org/10.3168/jds.2009-2864
  120. Xue GP, Johnson JS, Dalrymple BP. 1999. High osmolarity improves the electro-transformation efficiency of the grampositive bacteria Bacillus subtilis and Bacillus licheniformis. J. Microbiol. Methods 34: 183-191. https://doi.org/10.1016/S0167-7012(98)00087-6
  121. Yamane K, Bunai K, Kakeshita H. 2004. Protein traffic for secretion and related machinery of Bacillus subtilis. Biosci. Biotechnol. Biochem. 68: 2007-2023. https://doi.org/10.1271/bbb.68.2007
  122. Yang C, Song C, Freudl R, Mulchandani A, Qiao C. 2010. Twin-arginine translocation of methyl parathion hydrolase in Bacillus subtilis. Environ. Sci. Technol. 44: 7607-7612. https://doi.org/10.1021/es100860k
  123. Yang M, Zhang W, Ji S, Cao P, Chen Y, Zhao X. 2013. Generation of an artificial double promoter for protein expression in Bacillus subtilis through a promoter trap system. PLoS One 8: e56321. https://doi.org/10.1371/journal.pone.0056321
  124. Zanen G, Antelmann H, Westers H, Hecker M, van Dijl JM, Quax WJ. 2004. FlhF, the third signal recognition particle-GTPase of Bacillus subtilis, is dispensable for protein secretion. J. Bacteriol. 186: 5956-5960. https://doi.org/10.1128/JB.186.17.5956-5960.2004
  125. Zanen G, Houben EN, Meima R, Tjalsma H, Jongbloed JD, Westers H, et al. 2005. Signal peptide hydrophobicity is critical for early stages in protein export by Bacillus subtilis. FEBS J. 272: 4617-4630. https://doi.org/10.1111/j.1742-4658.2005.04777.x
  126. Zhang AL, Liu H, Yang MM, Gong YS, Chen H. 2007. Assay and characterization of a strong promoter element from B. subtilis. Biochem. Biophys. Res. Commun. 354: 90-95. https://doi.org/10.1016/j.bbrc.2006.12.137
  127. Zhang D, Sweredoski MJ, Graham RL, Hess S, Shan SO. 2012. Novel proteomic tools reveal essential roles of SRP and importance of proper membrane protein biogenesis. Mol. Cell Proteomics 11: M111.011585. https://doi.org/10.1074/mcp.M111.016345
  128. Zhang GQ, Bao P, Zhang Y, Deng AH, Chen N, Wen TY. 2011. Enhancing electro-transformation competency of recalcitrant Bacillus amyloliquefaciens by combining cell-wall weakening and cell-membrane fluidity disturbing. Anal. Biochem. 409: 130-137. https://doi.org/10.1016/j.ab.2010.10.013
  129. Zhang WW, Gao QR, Yang MM, Liu H, Wang D. 2012. Assay and characterization of an osmolarity inducible promoter newly isolated from Bacillus subtilis. Mol. Biol. Rep. 39: 7347-7353. https://doi.org/10.1007/s11033-012-1566-3
  130. Zhang XZ, You C, Zhang YH. 2014. Transformation of Bacillus subtilis. Methods Mol. Biol. 1151: 95-101. https://doi.org/10.1007/978-1-4939-0554-6_7
  131. Zimmer J, Li W, Rapoport TA. 2006. A novel dimer interface and conformational changes revealed by an X-ray structure of B. subtilis SecA. J. Mol. Biol. 364: 259-265. https://doi.org/10.1016/j.jmb.2006.08.044
  132. Zweers JC, Barak I, Becher D, Driessen AJ, Hecker M, Kontinen VP, et al. 2008. Towards the development of Bacillus subtilis as a cell factory for membrane proteins and protein complexes. Microb. Cell Fact. 7: 10. https://doi.org/10.1186/1475-2859-7-10

Cited by

  1. Efficient Secretion of the β-Galactosidase Bgal1-3 via both Tat-Dependent and Tat-Independent Pathways in Bacillus subtilis vol.64, pp.28, 2015, https://doi.org/10.1021/acs.jafc.6b01735
  2. Parametric continuous feed stream design to fine‐tune fed‐batch bioreactor performance: recombinant human growth hormone production in Bacillus subtilis vol.91, pp.11, 2015, https://doi.org/10.1002/jctb.4864
  3. Screening of Heterologous Signal Peptides for Optimization of the LIKE-Expression System vol.7, pp.2, 2015, https://doi.org/10.1007/s12668-016-0357-z
  4. Screening of Heterologous Signal Peptides for Optimization of the LIKE-Expression System vol.7, pp.2, 2015, https://doi.org/10.1007/s12668-016-0357-z
  5. High-Efficiency Secretion of β-Mannanase in Bacillus subtilis through Protein Synthesis and Secretion Optimization vol.65, pp.12, 2017, https://doi.org/10.1021/acs.jafc.6b05528
  6. Protein secretion inCorynebacterium glutamicum vol.37, pp.4, 2015, https://doi.org/10.1080/07388551.2016.1206059
  7. A novel strategy to improve protein secretion via overexpression of the SppA signal peptide peptidase in Bacillus licheniformis vol.16, pp.None, 2015, https://doi.org/10.1186/s12934-017-0688-7
  8. A food-grade expression system for D -psicose 3-epimerase production in Bacillus subtilis using an alanine racemase-encoding selection marker vol.4, pp.None, 2015, https://doi.org/10.1186/s40643-017-0139-7
  9. Heterologous signal peptides-directing secretion of Streptomyces mobaraensis transglutaminase by Bacillus subtilis vol.102, pp.13, 2018, https://doi.org/10.1007/s00253-018-9000-y
  10. Boosting heterologous protein production yield by adjusting global nitrogen and carbon metabolic regulatory networks in Bacillus subtilis vol.49, pp.None, 2015, https://doi.org/10.1016/j.ymben.2018.08.001
  11. Bacillus subtilis highly efficient protein expression systems that are chromosomally integrated and controllable by glucose and rhamnose vol.82, pp.11, 2018, https://doi.org/10.1080/09168451.2018.1497945
  12. Genetic and substrate-level modulation of Bacillus subtilis physiology for enhanced extracellular human interferon gamma production vol.48, pp.5, 2015, https://doi.org/10.1080/10826068.2018.1446157
  13. Unraveling Gardnerella vaginalis Surface Proteins Using Cell Shaving Proteomics vol.9, pp.None, 2015, https://doi.org/10.3389/fmicb.2018.00975
  14. Genetic Tools and Techniques for Recombinant Expression in Thermophilic Bacillaceae vol.6, pp.2, 2015, https://doi.org/10.3390/microorganisms6020042
  15. Core element characterization of Rhodococcus promoters and development of a promoter-RBS mini-pool with different activity levels for efficient gene expression vol.44, pp.None, 2015, https://doi.org/10.1016/j.nbt.2018.04.005
  16. Evaluation of promoter sequences for the secretory production of a Clostridium thermocellum cellulase in Paenibacillus polymyxa vol.102, pp.23, 2015, https://doi.org/10.1007/s00253-018-9369-7
  17. The Goldilocks Approach: A Review of Employing Design of Experiments in Prokaryotic Recombinant Protein Production vol.5, pp.4, 2015, https://doi.org/10.3390/bioengineering5040089
  18. Systematic Screening of Optimal Signal Peptides for Secretory Production of Heterologous Proteins in Bacillus subtilis vol.66, pp.50, 2018, https://doi.org/10.1021/acs.jafc.8b04183
  19. Tat-Independent Secretion of Polyethylene Terephthalate Hydrolase PETase in Bacillus subtilis 168 Mediated by Its Native Signal Peptide vol.66, pp.50, 2015, https://doi.org/10.1021/acs.jafc.8b05038
  20. 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
  21. Factors Influencing Recombinant Protein Secretion Efficiency in Gram-Positive Bacteria: Signal Peptide and Beyond vol.7, pp.None, 2015, https://doi.org/10.3389/fbioe.2019.00139
  22. Translation enhancement by a Dictyostelium gene sequence in Escherichia coli vol.103, pp.8, 2015, https://doi.org/10.1007/s00253-019-09746-7
  23. EngineeringBacillusfor efficient production of heterologous protein: current progress, challenge and prospect vol.126, pp.6, 2019, https://doi.org/10.1111/jam.14192
  24. Enhanced extracellular recombinant keratinase activity in Bacillus subtilis SCK6 through signal peptide optimization and site-directed mutagenesis vol.9, pp.57, 2015, https://doi.org/10.1039/c9ra07866e
  25. High-efficiency expression and secretion of human FGF21 in Bacillus subtilis by intercalation of a mini-cistron cassette and combinatorial optimization of cell regulatory components vol.18, pp.None, 2015, https://doi.org/10.1186/s12934-019-1066-4
  26. Cell Membrane and Electron Transfer Engineering for Improved Synthesis of Menaquinone-7 in Bacillus subtilis vol.23, pp.3, 2015, https://doi.org/10.1016/j.isci.2020.100918
  27. Systematic engineering of transport and transcription to boost alkaline α-amylase production in Bacillus subtilis vol.104, pp.7, 2015, https://doi.org/10.1007/s00253-020-10435-z
  28. Strategies for recombinant production of antimicrobial peptides with pharmacological potential vol.13, pp.4, 2020, https://doi.org/10.1080/17512433.2020.1764347
  29. Efficient secretory production of large‐size heterologous enzymes in Bacillus subtilis: A secretory partner and directed evolution vol.117, pp.10, 2015, https://doi.org/10.1002/bit.27478
  30. Bacillus subtilis : a universal cell factory for industry, agriculture, biomaterials and medicine vol.19, pp.None, 2015, https://doi.org/10.1186/s12934-020-01436-8
  31. Engineering a new vaccine platform for heterologous antigen delivery in live-attenuated Mycobacterium tuberculosis vol.19, pp.None, 2021, https://doi.org/10.1016/j.csbj.2021.07.035
  32. Construction of a Super-Folder Fluorescent Protein-Guided Secretory Expression System for the Production of Phospholipase D in Bacillus subtilis vol.69, pp.24, 2021, https://doi.org/10.1021/acs.jafc.1c02089
  33. An overview and future prospects of recombinant protein production in Bacillus subtilis vol.105, pp.18, 2015, https://doi.org/10.1007/s00253-021-11533-2
  34. Engineering Bacillus subtilis for the formation of a durable living biocomposite material vol.12, pp.1, 2015, https://doi.org/10.1038/s41467-021-27467-2
  35. Production of d-Tagatose by Whole-Cell Conversion of Recombinant Bacillus subtilis in the Absence of Antibiotics vol.10, pp.12, 2015, https://doi.org/10.3390/biology10121343
  36. Metabolic engineering of Bacillus subtilis with an endopolygalacturonase gene isolated from Pectobacterium. carotovorum; a plant pathogenic bacterial strain vol.16, pp.12, 2021, https://doi.org/10.1371/journal.pone.0256562