Next Article in Journal
Prevalence of Infectious Agents Causing Abortion in Pregnant Women Using Serological Tests and Histopathological Analysis
Next Article in Special Issue
The Effect of Antibiotic Treatment and Gene Expression of Mex B Efflux Transporters on the Resistance in Pseudomonas Aeruginosa Biofilms
Previous Article in Journal
A Non-Electrolysis Bioelectric Effect for Gingivitis and Hygiene Contamination Biofilm Removal
Previous Article in Special Issue
Pterostilbene, a Natural Methoxylated Analog of Resveratrol, Exhibits Antifungal Activity Induced by Reactive Oxygen Species Production and Plasma Membrane Injury
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Translation Enhancement by a Short Nucleotide Insertion at 5′UTR: Application to an In Vitro Cell-Free System and a Photosynthetic Bacterium

Graduate School of Arts and Sciences, The University of Tokyo, 3-8-1 Komaba, Meguro-ku, Tokyo 153-8902, Japan
*
Author to whom correspondence should be addressed.
Appl. Microbiol. 2023, 3(3), 687-697; https://doi.org/10.3390/applmicrobiol3030047
Submission received: 12 June 2023 / Revised: 27 June 2023 / Accepted: 28 June 2023 / Published: 2 July 2023

Abstract

:
We previously showed that insertion of Dictyostelium gene sequences, such as mlcR, upstream of the Shine–Dalgarno sequence, positively impacts downstream gene expression in Escherichia coli. However, the mechanism by which protein production is facilitated and its applicability to other bacteria remains unknown. In this study, a translation-enhancing effect, associated with this system, on the mRNA amount and property as well as the versatility of the method has been demonstrated. The insertion of mlcR-terminal 25 bp (mlcR25) stabilized the mRNAs and led to increased mRNA levels in E. coli. In the in vitro translation system, a four-fold enhancement was observed when DNA was used as the template, and a three-fold enhancement was observed when mRNA was used as the template. This suggests that mlcR25 has an effect on the facilitation of the interaction between mRNA and ribosome. Furthermore, when this enhancement system was adapted to the photosynthetic bacterium Rhodobacter capsulatus, a more than six-fold increase in translation was observed. Thus, we propose that enhanced translation by mlcR25 is mediated by mechanisms that help the translation machinery to work efficiently, and the system can be applied to bacteria other than E. coli.

Graphical Abstract

1. Introduction

Gene expression involves the transcription of mRNA and its translation into proteins. Escherichia coli, a Gram-negative γ-proteobacterium, is a model organism for which the mechanism of gene expression has been studied extensively. Important factors that affect transcription include sequence characteristics, such as the promoter, terminator, Shine–Dalgarno (SD) sequence [1], and ribosome-binding region in the 5′- or 3′-untranslated region (UTR). In addition, recent accumulating evidence indicates that mRNA stability and codon usage are important factors for the regulation of gene expression [2,3,4,5,6,7,8,9,10,11]. Escherichia coli is suitable for producing various functional proteins at a low cost and high yield; thus, methods are being developed to regulate the protein-expression levels [12].
The strength of the aforementioned transcriptional factors varies owing to differences in their sequences. Therefore, while the simplest approach is to identify a promoter with enhanced activation, attempts are also being made to optimize the upstream sequence of the SD sequence [12]. Recently, we reported that the insertion of a gene sequence from the eukaryotic cellular slime mold Dictyostelium discoideum (e.g., mlcR encoding myosin regulatory light chain) upstream of the SD sequence increased protein production in E. coli [13]. We named this phenomenon Translation Enhancement by a Dictyostelium gene sequence (TED). TED is a straightforward method used to increase protein expression levels by inserting a short sequence in the 5′-UTR (Figure 1). In this method, the insertion or substitution of the Dictyostelium gene sequence into an existing vector or possibly the genome enhances protein synthesis, which leads to an increased yield of the desired protein. Currently, TED that uses a 25 bp sequence of the 3′-end of mlcR from Dictyostelium (mlcR25) has been the most effective [13]. Replacing the T7 phi10 [14] in the pET vector, a typical vector for protein production, with mlcR25 can further increase production. An interesting aspect of TED used for mlcR25 is that sufficient fluorescence emitted from the green fluorescent protein (GFP), with low levels of transcription leakage from the lac promoter, was visually observed even in E. coli cells [13]. This suggests that TED may have a positive effect on translation rather than transcription. However, the mechanism underlying this effect remains unclear.
The insertion of a short sequence may enhance downstream gene expression in other bacteria; however, this is still unknown. Typically, the 20–50 a.a.-length short leader peptides were employed for a bicistronic design expression system [15,16,17]. In this system, a short peptide (the first cistron) is generally translated using the first SD sequence in the 5′UTR and a second SD site containing the first cistron initiates the translation of the target gene, which is usually called the second cistron, under the control of a single promoter [18,19]. In the case of several fungi, the insertion of a ~100 bp fragment containing the cis-element into a heterologous promoter substantially accelerates the expression of the downstream gene [20,21].
In order to elucidate the underlying mechanism of mlcR25-based TED, we analyzed whether the levels of mRNAs containing mlcR25 and those without mlcR25 are altered in vivo and investigated whether properties of the mRNA itself promote translation using in vitro translation experiments. Because TED has only been observed intracellularly, the components of the cell utilized for TED and whether the phenomenon is dependent on the genetic background of the cell are unclear. Hence, we aimed to reproduce TED in vitro using a reconstituted cell-free protein synthesis system (the PURE system) [22]. This system is ideal for quantitative analysis because proteins can be synthesized using chemically defined components, and both DNA and mRNA can be used as a template. Moreover, we tested the availability of TED in the purple photosynthetic α-proteobacterium Rhodobacter capsulatus, which has long been studied in the field of photosynthesis and is widely used as a host for the expression of recombinant proteins to understand the utility of the TED system in different bacteria. We hypothesize that TED leads to increased protein production through translation control of mRNA in various bacteria.

2. Materials and Methods

2.1. Cell Culture

Escherichia coli K-12 derivatives DH5α and HST08 (Takara Bio, Shiga, Japan), which are widely used for molecular cloning, and XL1-Blue/pDPT51, which is used for transformation via conjugation in several bacteria, were cultured at 37 °C in lysogeny broth (LB) medium in the presence of appropriate antibiotics [23]. Rifampicin (FUJIFILM Wako Pure Chemical Corporation, Osaka, Japan) was dissolved in ethanol at a concentration of 2 mg/mL, and a final concentration of 2 µg/mL was used for growing the cells. Trimethoprim and spectinomycin were used at a concentration of 50 µg/mL and 40 µg/mL, respectively.
Rhodobacter capsulatus SB1003 was cultured aerobically at 30 °C in PYS [24], a rich medium, or RCV minimum medium [25]. Gentamycin, rifampicin, and spectinomycin were used at a concentration of 1.5 µg/mL, 100 µg/mL, and 10 µg/mL, respectively.

2.2. Plasmids

The insertion of mlcR25-RBS-GFP (#154295; Addgene) or RBS-GFP into the pUC19 vector has been described previously [13]. mlcR25-RBS-GFP or RBS-GFP was inserted into the pET vector using restriction enzyme digestion with enzymes XbaI (Takara Bio, Kusatsu, Japan) and BamHI (Takara Bio, Kusatsu, Japan), followed by ligation.
The sqr promoter and lacZ fusion plasmids were constructed using the pNM481 plasmid [26] with the Ω-interposon (Smr/Spr) gene [27] inserted upstream of the cloned fragment, which were transferred into R. capsulatus with the conjugative E. coli strain XL1-Blue/pDPT51, as described previously [28,29]. The insertion of sqr promoter into the pNM481 plasmid with the Ω-interposon gene has been described previously [30]. The insertion of mlcR25 into the sqr promoter region was accomplished by amplification of the sqr-promoter-inserted pNM481 without the Ω-interposon gene (pNM481:psqr) by PCR using KOD One with a mlcR25 containing primers (5′-cgttaatactctcttcagtaaaaaaggagggacagatggctcatatcgcc-3′) and (5′-tactgaagagagtattaacgaaaagccgaactggctgtcgggccgaagcc-3′) and circularization of the amplified DNA fragment by In-Fusion HD Cloning kit (Takara Bio, Kusatsu, Japan). The Ω-interposon (Smr/Spr) gene was then inserted into the plasmid at the SmaI site by ligation. Transformation of E. coli was performed chemically or by electroporation using the Gene Pulser Xcell system (Bio-Rad Laboratories, Hercules, CA, USA).

2.3. In Vitro Transcription

Template plasmid DNA was prepared using the QIAGEN Plasmid Midi Kit (QIAGEN, Hilden, Germany). After linearization using EcoRI, DNA was purified using phenol-chloroform extraction and ethanol precipitation. The T7 RiboMAX Large-Scale RNA Production System (Promega, Madison, WI, USA) was used for in vitro transcription. The synthesized mRNAs were purified using phenol extraction, followed by isopropanol precipitation, and were stored at −80 °C until further use.

2.4. In Vitro Translation

PUREflex 1.0 (GeneFrontier Corporation, Tokyo, Japan) was used as a cell-free protein expression system according to the manufacturer’s protocol. The composition of the system has been described previously [22,31]. A total of 10 µL of reaction mixture containing 5 µL of solution I, 0.5 µL of solution II, 1 µL of solution III, 2.5 µL of nuclease-free water, and 1 µL of DNA (180 ng) or mRNA (1 ng) was incubated at 37 °C for 6 h. Next, 10 µL of nuclease-free water and 20 µL of 3× Laemmli sample buffer were added to the reaction mixture to stop the reaction. The samples were stored at −20 °C until further use.

2.5. Western Blotting

The samples (3 µL) were subjected to SDS-PAGE and transferred to a polyvinylidene difluoride membrane (0.2 μm pore size, Bio-Rad Laboratories, Hercules, CA, USA). The membrane was blocked using 3% skim milk in PBS (137 mM NaCl, 2.7 mM KCl, 4 mM Na2HPO4·12H2O, and 0.7 mM KH2PO4; pH 7.4) for 1 h at 25 °C. After washing three times with PBS containing 0.05% Tween-20 (PBS-T), the membrane was incubated with rabbit anti-GFP antibody (MBL, Tokyo, Japan; Code No. 598) at 1:1000 dilution with Can Get Signal solution 1 (TOYOBO, Osaka, Japan) for 1.5 h at 25 °C. After washing with PBS-T three times, the membrane was incubated with horseradish peroxidase-conjugated anti-rabbit IgG (GE Healthcare, Chicago, IL, USA) at 1:5000 dilution with Can Get Signal solution 2 (TOYOBO) for 1 h at 25 °C. Blots were visualized using the enhanced chemiluminescence procedure with the Immobilon Western chemiluminescent HRP substrate (Millipore Corporation, Billerica, MA, USA) and imaged using an ImageQuant LAS 4000 (GE Healthcare, Chicago, IL, USA). A standard CBB solution or Quick-CBB PLUS solution (FUJIFILM Wako Pure Chemical Corporation, Osaka, Japan) was used to stain the transferred proteins on the membrane. Band densitometry analysis was performed using the Fiji/ImageJ software [32]. The value for the GFP band was normalized to that of the control protein (~45 kDa) band observed in the same lane stained with CBB.

2.6. RNA Purification and Real-Time PCR

After culturing the cells in LB medium containing antibiotics, 7 × 108 cells in the logarithmic growth phase were collected by centrifugation and treated with lysozyme for 10 min at 37 °C. RNA was then purified using NucleoSpin RNA Plus (Takara Bio), according to the manufacturer’s protocol. The RNA concentration was measured using NanoDrop (Thermo Fisher Scientific, Carlsbad, CA, USA). Then, 100 ng of RNA in the mixture was used for reverse transcription using the PrimeScript RT reagent kit (Takara Bio). Real-time PCR was performed with TB Green Premix Ex Taq II (Tli RNaseH Plus) (Takara Bio) in a 25 µL volume using a Thermal Cycler Dice Real Time System III (Takara Bio). Two cDNA dilutions (1:2 and 1:4) were used, and duplicates were set for each reaction. The extent of amplification was evaluated using the 2−ΔΔCt method. The cysG was used as an internal control to normalize the cDNA input [33]. The following primers were used: gfp Forward: 5′-ggtgaaggtgaaggagatgc-3′ and Reverse: 5′-taggccagggtacaggtaac-3′, and cysG Forward: 5′-attgaacacggaatgccagg-3′ and Reverse: 5′-gtgagcgtaccgtcaatcac-3′.

2.7. β-Galactosidase Assay

Rhodobacter capsulatus cells containing the sqr promoter region and lacZ fusion plasmid were grown aerobically to the mid-log phase in RCV medium. For sulfide induction, a final measure of 0.6 mM of Na2S was added and cells were grown further for 120 min. After the induction, 15 mL of cells were harvested, and β-galactosidase activity was determined essentially as described previously [28]. The results were obtained as the amount of o-nitrophenyl-β-D-galactopyranoside (ONPG) hydrolyzed per min per mg of protein.

2.8. Statistical Analysis

Statistical analyses were performed using GraphPad Prism9 (GraphPad Software, La Jolla, CA, USA). The unpaired t-test, one-way ANOVA, and post hoc Tukey’s multiple comparisons test were performed as indicated. Normality was tested using the Shapiro–Wilk test or the Kolmogorov–Smirnov test. A p-value < 0.05 was considered statistically significant.

3. Results

3.1. Effect of mlcR25 on mRNA Longevity in E. coli

It has been demonstrated that the AU-rich sequence in the 5′-UTR stabilizes the mRNA [34]. No correlation between AT% and its enhancing effect has been found in TED [13]; however, since mlcR25 comprises 72% of the AT-rich sequence (Figure 1b), the longevity of mRNA in cells harboring the plasmids with or without mlcR25 under the control of the lac promoter (Figure 2a) was investigated [13]. Rifampicin, an inhibitor of the bacterial DNA-dependent RNA polymerase [35], was used to inhibit transcription. After rifampicin treatment, the total RNA was isolated from the cells at designated time points. The total amount of gfp mRNA without mlcR25 (RBS-GFP) was found to have decreased over time, whereas that of the gfp mRNAs with mlcR25 (mlcR25-RBS-GFP) was constant, suggesting that mRNAs containing mlcR25 remained stable for a longer duration than those without (Figure 2b). Finally, the amount of gfp mRNA in the cells containing each plasmid was compared. Consistent with the increased longevity of gfp mRNA, the gfp mRNAs with mlcR25 were found to be, on average, four times more abundant than those without mlcR25 (Figure 2c). Thus, mlcR25 had a positive impact on mRNA longevity. Since a high mRNA content is assumed to increase the frequency of contact with the ribosome, the mRNA content is one possible cause of TED.

3.2. mlcR25 Confers Enhanced Translational Efficiency on mRNA

An in vitro translation system was used to examine factors other than mRNA levels. A plasmid containing a ribosome-binding site (RBS) and inserted mlcR25 to express GFP under the control of the T7 promoter to use in the PURE system [22,31] (Figure 3a) was constructed. The PURE system includes the T7 RNA polymerase and ribosomes to perform a series of mRNA and protein syntheses. First, DNA was used as a template to test whether TED works in the system. GFP (~27 kDa) synthesis from plasmids containing mlcR25 was detected by Western blotting; however, that from plasmids without mlcR25 was at a very low level (Figure 3b). The GFP band density was, on average, four times higher for mlcR25-RBS-GFP than for RBS-GFP (Figure 3c). These data demonstrate that TED can be reproduced in vitro, suggesting that it is driven by factors included in the PURE system.
Next, to examine whether mlcR25 promotes translation at the mRNA level, mRNA was used as a template in the PURE system. In the above experiment, owing to the addition of equal amounts of DNA to the reaction mixtures, equal amounts of GFP mRNA synthesized by T7 RNA polymerase were assumed to be present in the reaction mixture. The PURE system used does not contain mRNA-degrading enzymes, such as RNase E; thus, if mlcR25 affects translation rather than transcription, even with the addition of equal amounts of mRNA in the system, an increase in the translation of mlcR25-RBS-GFP mRNA compared to that of RBS-GFP would be expected. An in vitro translation using mRNA instead of DNA was used to verify this hypothesis. The mRNA containing mlcR25 showed, on average, three-times-higher protein synthesis than an equal amount of mRNA without mlcR25 (Figure 3d,e). Hence, it was concluded that TED using mlcR25 leads to increased protein expression through mRNA regulation at the translational level.

3.3. Application to a Photosynthetic Bacterium

To examine the effect of mlcR25 in other bacteria, we used R. capsulatus with a lacZ reporter system driven by the sqr promoter (Figure 4a). In this construct, mlcR25 was inserted upstream of sqr encoding sulfide-quinone reductase (SQR) and further fused with lacZ. The sqr promoter is repressed by sulfide-responsive transcription factor SqrR in the absence of sulfide [30]. Therefore, the sqr promoter driving lacZ expression was induced by treating the cells with sulfide, and the translation level of lacZ was subsequently measured as its enzymatic activity. Importantly, we found that the insertion of mlcR25 upstream of the SD sequence clearly increased the activity of β-galactosidase by about sixfold (Figure 4c). This suggests that TED functions in other bacteria as well as in E. coli.

4. Discussion

In this study, it was observed that mlcR25 extended the lifespan of mRNA in E. coli cells. We demonstrated that the insertion of mlcR25 upstream of RBS promotes translation in vitro, regardless of DNA or RNA being used as the template, similar to the effect observed in cells. Furthermore, the translation-enhancing effect of mlcR25 was also observed in R. capsulatus, a bacterium from a different class.
For in vitro translation, we used the PURE system [22,31], which consists of highly purified proteins required for transcription, translation, aminoacylation, and energy regeneration as well as amino acids and NTPs and demonstrated that mlcR25-mediated TED is facilitated during the translation process in the system. The following components are included in this system: IF1/IF2/IF3 as initiation factors, EF-Tu/EF-Ts/EF-G as elongation factors, RF1/RF2/RF3 as release factors, ribosome recycling factor, 20 kinds of aminoacyl-tRNA synthetase, and methionyl-tRNA transformylase. The fact that the components necessary for transcription–translation are known is an advantage because it eliminates unknown material in the cell. Hence, we deemed this system to be appropriate for the purpose of this study.
Cell-free systems such as the PURE system have been used to synthesize membrane proteins, which are generally difficult to obtain, and antibodies for mRNA/ribosome display [36,37,38,39]. The PURE system is expected to be a fundamental method for synthetic biology [40,41], and improvements in product solubility and synthetic cost are still being made [42,43]. We suggest that the method using mlcR25 presented in this paper is useful because it can be immediately adapted to regulate the expression efficiency.
We used R. capsulatus as a model case for the TED-adapted non-E. coli protein expression system. This bacterium develops an intracytoplasmic membrane when growing photosynthetically in order to perform photosynthesis effectively by increasing the surface area of the membrane; therefore, R. capsulatus and phylogenetically closely related R. sphaeroides are utilized for the overexpression and purification of membrane proteins [44,45,46]. Rhodobacter capsulatus is also used for the functional expression of cofactor-dependent enzymes because it can produce a wider variety of metal-containing coenzymes as compared with E. coli [47]. Several vectors have also been constructed to optimize and control the expression of these various proteins [48]. Therefore, the availability of TED in this bacterium has significant bioengineering potential. Taxonomically, E. coli is a γ-proteobacteria, whereas R. capsulatus belongs to the class α-proteobacteria. The transcriptional mechanisms of RNAP/σ70 of E. coli and R. capsulatus are likely similar because they are structurally and functionally similar in part [49]. In translation, the SD sequences are conserved in the genes of R. capsulatus and R. sphaeroides [50,51] as in other bacteria [52]. Although there are differences between the ribosomes of E. coli and R. capsulatus [53,54], we assume that there are similarities in the translation mechanism, at least at the initiation stage. Thus, TED may be adaptable to the broad bacteria.
It remains unclear how the insertion of mlcR25 in the 5′-UTR helps in the ribosome function. Komarova et al. (2005, [34]) reported that AU-rich sequences promote the binding of ribosomal protein S1. This protein is known to contribute to mRNA unfolding during translation initiation [55]. Furthermore, translation is suppressed when there is a stable secondary structure in the RBS [56]. Therefore, AU-rich sequences should be placed in this region to reduce the stability of mRNA and increase its acceptability by ribosomal protein S1. Interestingly, mlcR25 has been predicted to form a secondary structure with neighboring sequences [13]. The relationship between mlcR25 and ribosomal protein S1 should be investigated in the future.
The reason for the stability of mRNAs containing mlcR25 for a long duration is unclear. One possibility is that they are stabilized owing to the formation of a stem-loop structure that is less susceptible to degradation, as observed in studies pertaining to ompA [57,58,59]. The deletion of a 104 bp sequence, including the region that forms the stem loom upstream of the SD of innate ompA, promotes mRNA degradation. Importantly, the insertion of synthetic sequences with 5′ self-complementarity to create a stem loop restores the stability. Thus, based on our prediction of the stem loop created by mlcR25 [13], such secondary structures may contribute to the observed mRNA stabilization. Another possible mechanism for the longevity of the mRNAs containing mlcR25 is the coupling of transcription and translation. In bacteria, transcription and translation occur simultaneously (i.e., the ribosome binds and translation begins before transcription is completed) [3,7,60]. One scenario is that the transcribed mlcR25-containing mRNA is immediately bound to the ribosome and translated, which may result in protection from RNase and increase its net mRNA content. mRNAs that are translationally inefficient are actively degraded [61]. On the contrary, highly translated mRNAs are less likely to be degraded [2,9,62]. Similarly, a previous study on ompA has also indicated that ribosome binding at the 5′-UTR induced mRNA stability [58]. Considering these findings, an increased stability and level of mRNAs with mlcR25 supports the fact that translation is promoted by mlcR25.

5. Conclusions

In conclusion, TED with mlcR25 promotes translation by acting as a cis-acting factor in E. coli, R. capsulatus, and the PURE system. The fact that translation was promoted by mRNA in the PURE system in this study suggests that mlcR25 promotes the interaction of mRNAs containing this sequence with ribosomes. Additionally, the amount of mRNA containing mlcR25 was also increased, which also increases the frequency of contact between mRNA and ribosome, and thus may help promote the translation of downstream genes. In addition to the improvement of the gene structure shown in this study, further improvement of the transcription apparatus and ribosome is also possible at the genomic level, and tuning both elements should be emphasized in the future.

Author Contributions

Conceptualization, T.K.; methodology, T.K. and T.S.; validation, T.K. and T.S.; formal analysis, T.K. and T.S.; investigation, T.K. and T.S.; resources, T.K. and T.S.; data curation, T.K. and T.S.; writing—original draft preparation, T.K.; writing—review and editing, T.K. and T.S.; visualization, T.K. and T.S.; supervision, T.K.; project administration, T.K.; funding acquisition, T.K. and T.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Japan Society for the Promotion of Science KAKENHI Grant Number 19K15809 to T.K., 21K15038 and 21H05271 to T.S.

Data Availability Statement

Not applicable.

Acknowledgments

We thank Miho Ohsugi and Tatsuru Masuda for providing the experimental equipment, and Kohtoh Yukawa and Tomomi Taniguchi for helping with the experiments in the early stages of the research.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Shine, J.; Dalgarno, L. The 3’ Terminal Sequence of Escherichia coli 16S Ribosomal RNA: Complementarity to Nonsense Triplets and Ribosome Binding Sites. Proc. Natl. Acad. Sci. USA 1974, 71, 1342–1346. [Google Scholar] [CrossRef]
  2. Boël, G.; Letso, R.; Neely, H.; Price, W.N.; Wong, K.H.; Su, M.; Luff, J.D.; Valecha, M.; Everett, J.K.; Acton, T.B.; et al. Codon Influence on Protein Expression in E. coli Correlates with mRNA Levels. Nature 2016, 529, 358–363. [Google Scholar] [CrossRef] [Green Version]
  3. Belasco, J.G. mRNA Degradation in Prokaryotic Cells: An Overview; Woodhead Publishing Limited: Sawston, UK, 1993. [Google Scholar]
  4. Carrier, T.A.; Keasling, J.D. Controlling Messenger RNA Stability in Bacteria: Strategies for Engineering Gene Expression. Biotechnol. Prog. 1997, 13, 699–708. [Google Scholar] [CrossRef] [PubMed]
  5. Carrier, T.A.; Keasling, J.D. Engineering mRNA Stability in E. coli by the Addition of Synthetic Hairpins Using a 5′ Cassette System. Biotechnol. Bioeng. 1997, 55, 577–580. [Google Scholar] [CrossRef]
  6. Carrier, T.A.; Keasling, J.D. Library of Synthetic 5’ Secondary Structures To Manipulate mRNA Stability in Escherichia coli. Biotechnol. Prog. 1999, 15, 58–64. [Google Scholar] [CrossRef]
  7. Radhakrishnan, A.; Green, R. Connections Underlying Translation and mRNA Stability. J. Mol. Biol. 2016, 428, 3558–3564. [Google Scholar] [CrossRef] [PubMed]
  8. Grunberg-Manago, M. Messenger RNA Stability and Its Role in Control of Gene Expression in Bacteria and Phages. Annu. Rev. Genet. 1999, 33, 193–227. [Google Scholar] [CrossRef]
  9. Kudla, G.; Murray, A.W.; Tollervey, D.; Plotkin, J.B. Coding-Sequence Determinants of Gene Expression in Escherichia coli. Science 2009, 324, 255–258. [Google Scholar] [CrossRef] [Green Version]
  10. Plotkin, J.B.; Kudla, G. Synonymous but Not the Same: The Causes and Consequences of Codon Bias. Nat. Rev. Genet. 2011, 12, 32–42. [Google Scholar] [CrossRef] [Green Version]
  11. Hanson, G.; Coller, J. Translation and Protein Quality Control: Codon Optimality, Bias and Usage in Translation and mRNA Decay. Nat. Rev. Mol. Cell. Biol. 2018, 19, 20–30. [Google Scholar] [CrossRef]
  12. Kondo, T.; Yumura, S. Strategies for Enhancing Gene Expression in Escherichia coli. Appl. Microbiol. Biotechnol. 2020, 104, 3825–3834. [Google Scholar] [CrossRef] [PubMed]
  13. Kondo, T.; Yumura, S. Translation Enhancement by a Dictyostelium Gene Sequence in Escherichia coli. Appl. Microbiol. Biotechnol. 2019, 103, 3501–3510. [Google Scholar] [CrossRef] [PubMed]
  14. Olins, P.O.; Rangwala, S.H. A Novel Sequence Element Derived from Bacteriophage T7 mRNA Acts as an Enhancer of Translation of the LacZ Gene in Escherichia coli. J. Biol. Chem. 1989, 264, 16973–16976. [Google Scholar] [CrossRef]
  15. Jang, S.H.; Cha, J.W.; Han, N.S.; Jeong, K.J. Development of Bicistronic Expression System for the Enhanced and Reliable Production of Recombinant Proteins in Leuconostoc Citreum. Sci. Rep. 2018, 8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Sun, M.; Gao, X.; Zhao, Z.; Li, A.; Wang, Y.; Yang, Y.; Liu, X.; Bai, Z. Enhanced Production of Recombinant Proteins in Corynebacterium Glutamicum by Constructing a Bicistronic Gene Expression System. Microb. Cell. Fact. 2020, 19, 113. [Google Scholar] [CrossRef]
  17. Li, D.; Fu, G.; Tu, R.; Jin, Z.; Zhang, D. 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. Microb. Cell. Fact. 2019, 18, 17. [Google Scholar] [CrossRef] [Green Version]
  18. Schoner, B.E.; Belagaje, R.M.; Schoner, R.G. Translation of a Synthetic Two-Cistron mRNA in Escherichia coli. Proc. Natl. Acad. Sci. USA 1986, 83, 8506–8510. [Google Scholar] [CrossRef]
  19. Mutalik, V.K.; Guimaraes, J.C.; Cambray, G.; Lam, C.; Christoffersen, M.J.; Mai, Q.-A.; Tran, A.B.; Paull, M.; Keasling, J.D.; Arkin, A.P.; et al. Precise and Reliable Gene Expression via Standard Transcription and Translation Initiation Elements. Nat. Methods 2013, 10, 354–360. [Google Scholar] [CrossRef]
  20. Ishida, H.; Hata, Y.; Kawato, A.; Abe, Y. Improvement of the GlaB Promoter Expressed in Solid-State Fermentation (SSF) of Aspergillus Oryzae. Biosci. Biotechnol. Biochem. 2006, 70, 1181–1187. [Google Scholar] [CrossRef] [Green Version]
  21. Deng, J.; Wu, Y.; Zheng, Z.; Chen, N.; Luo, X.; Tang, H.; Keasling, J.D. A Synthetic Promoter System for Well-Controlled Protein Expression with Different Carbon Sources in Saccharomyces cerevisiae. Microb. Cell. Fact. 2021, 20, 202. [Google Scholar] [CrossRef]
  22. Shimizu, Y.; Inoue, A.; Tomari, Y.; Suzuki, T.; Yokogawa, T.; Nishikawa, K.; Ueda, T. Cell-Free Translation Reconstituted with Purified Components. Nat. Biotechnol. 2001, 19, 751–755. [Google Scholar] [CrossRef] [PubMed]
  23. Kondo, T.; Yumura, S. An Improved Molecular Tool for Screening Bacterial Colonies Using GFP Expression Enhanced by a Dictyostelium Sequence. Biotechniques 2020, 68, 91–95. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Sato, T.; Inoue, K.; Sakurai, H.; Nagashima, K.V.P. Effects of the Deletion of Hup Genes Encoding the Uptake Hydrogenase on the Activity of Hydrogen Production in the Purple Photosynthetic Bacterium Rubrivivax gelatinosus IL144. J. Gen. Appl. Microbiol. 2017, 63, 274–279. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Weaver, P.F.; Wall, J.D.; Gest, H. Characterization of Rhodopseudomonas capsulata. Arch Microbiol. 1975, 105, 207–216. [Google Scholar] [CrossRef]
  26. Minton, N.P. Improved Plasmid Vectors for the Isolation of Translational Lac Gene Fusions. Gene 1984, 31, 269–273. [Google Scholar] [CrossRef]
  27. Fellay, R.; Frey, J.; Krisch, H. Interposon Mutagenesis of Soil and Water Bacteria: A Family of DNA Fragments Designed for in Vitro Insertional Mutagenesis of Gram-Negative Bacteria. Gene 1987, 52, 147–154. [Google Scholar] [CrossRef]
  28. Young, D.A.; Bauer, C.E.; Williams, J.C.; Marrs, B.L. Gentic Evidence for Superoperonal Organization of Genes for Photosynthesis Pigments and Pigment-Binding Proteins in Rhodobacter capsulatus. Mol. Gen. Genet. 1989, 218, 1–12. [Google Scholar] [CrossRef]
  29. Sganga, M.W.; Bauer, C.E. Regulatory Factors Controlling Photosynthetic Reaction Center and Light-Harvesting Gene Expression in Rhodobacter capsulatus. Cell 1992, 68, 945–954. [Google Scholar] [CrossRef]
  30. Shimizu, T.; Shen, J.; Fang, M.; Zhang, Y.; Hori, K.; Trinidad, J.C.; Bauer, C.E.; Giedroc, D.P.; Masuda, S. Sulfide-Responsive Transcriptional Repressor SqrR Functions as a Master Regulator of Sulfide-Dependent Photosynthesis. Proc. Natl. Acad. Sci. USA 2017, 114, 2355–2360. [Google Scholar] [CrossRef]
  31. Shimizu, Y.; Kanamori, T.; Ueda, T. Protein Synthesis by Pure Translation Systems. Methods 2005, 36, 299–304. [Google Scholar] [CrossRef]
  32. Schindelin, J.; Arganda-Carreras, I.; Frise, E.; Kaynig, V.; Longair, M.; Pietzsch, T.; Preibisch, S.; Rueden, C.; Saalfeld, S.; Schmid, B.; et al. Fiji: An Open-Source Platform for Biological-Image Analysis. Nat. Methods 2012, 9, 676–682. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Zhou, K.; Zhou, L.; Lim, Q.; Zou, R.; Stephanopoulos, G.; Too, H.P. Novel Reference Genes for Quantifying Transcriptional Responses of Escherichia coli to Protein Overexpression by Quantitative PCR. BMC Mol. Biol. 2011, 12, 1–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Komarova, A.V.; Tchufistova, L.S.; Dreyfus, M.; Boni, I.V. AU-Rich Sequences within 5’ Untranslated Leaders Enhance Translation and Stabilize mRNA in Escherichia coli. J. Bacteriol. 2005, 187, 1344–1349. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Umezawa, H.; Mizuno, S.; Yamazaki, H.; Nitta, K. Inhibition of DNA-Dependent RNA Synthesis by Rifamycins. J. Antibiot. (Tokyo) 1968, 21, 234–236. [Google Scholar] [CrossRef] [Green Version]
  36. Kanamori, T.; Fujino, Y.; Ueda, T. PURE Ribosome Display and Its Application in Antibody Technology. Biochim. Biophys. Acta. Proteins Proteom. 2014, 1844, 1925–1932. [Google Scholar] [CrossRef]
  37. Endo, Y.; Takai, K.; Ueda, T. Cell-Free Protein Production; Endo, Y., Takai, K., Ueda, T., Eds.; Methods in Molecular Biology; Humana Press: Totowa, NJ, USA, 2010; Volume 607, ISBN 978-1-60327-330-5. [Google Scholar]
  38. Nagumo, Y.; Fujiwara, K.; Horisawa, K.; Yanagawa, H.; Doi, N. PURE mRNA Display for in Vitro Selection of Single-Chain Antibodies. J. Biochem. 2016, 159, 519–526. [Google Scholar] [CrossRef]
  39. Katzen, F.; Fletcher, J.E.; Yang, J.P.; Kang, D.; Peterson, T.C.; Cappuccio, J.A.; Blanchette, C.D.; Sulchek, T.; Chromy, B.A.; Hoeprich, P.D.; et al. Insertion of Membrane Proteins into Discoidal Membranes Using a Cell-Free Protein Expression Approach. J. Proteome Res. 2008, 7, 3535–3542. [Google Scholar] [CrossRef]
  40. Forster, A.C.; Church, G.M. Towards Synthesis of a Minimal Cell. Mol. Syst. Biol. 2006, 2. [Google Scholar] [CrossRef] [Green Version]
  41. Matsubayashi, H.; Ueda, T. Purified Cell-Free Systems as Standard Parts for Synthetic Biology. Curr. Opin. Chem. Biol. 2014, 22, 158–162. [Google Scholar] [CrossRef]
  42. Niwa, T.; Kanamori, T.; Ueda, T.; Taguchi, H. Global Analysis of Chaperone Effects Using a Reconstituted Cell-Free Translation System. Proc. Natl. Acad. Sci. USA 2012, 109, 8937–8942. [Google Scholar] [CrossRef]
  43. Lavickova, B.; Maerkl, S.J. A Simple, Robust, and Low-Cost Method to Produce the PURE Cell-Free System. ACS Synth. Biol. 2019, 8, 455–462. [Google Scholar] [CrossRef]
  44. Roy, A.; Shukla, A.K.; Haase, W.; Michel, H. Employing Rhodobacter sphaeroides to Functionally Express and Purify Human G Protein-Coupled Receptors. Biol. Chem. 2008, 389, 69–78. [Google Scholar] [CrossRef] [PubMed]
  45. Erbakan, M.; Curtis, B.S.; Nixon, B.T.; Kumar, M.; Curtis, W.R. Advancing Rhodobacter sphaeroides as a Platform for Expression of Functional Membrane Proteins. Protein. Expr. Purif. 2015, 115, 109–117. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Laible, P.D.; Scott, H.N.; Henry, L.; Hanson, D.K. Towards Higher-Throughput Membrane Protein Production for Structural Genomics Initiatives. J. Struct. Funct. Genom. 2004, 5, 167–172. [Google Scholar] [CrossRef] [PubMed]
  47. Kappler, U.; McEwan, A.G. A System for the Heterologous Expression of Complex Redox Proteins in Rhodobacter capsulatus: Characterisation of Recombinant Sulphite:Cytochrome c Oxidoreductase from Starkeya Novella. FEBS Lett. 2002, 529, 208–214. [Google Scholar] [CrossRef] [Green Version]
  48. Katzke, N.; Arvani, S.; Bergmann, R.; Circolone, F.; Markert, A.; Svensson, V.; Jaeger, K.-E.; Heck, A.; Drepper, T. A Novel T7 RNA Polymerase Dependent Expression System for High-Level Protein Production in the Phototrophic Bacterium Rhodobacter capsulatus. Protein. Expr. Purif. 2010, 69, 137–146. [Google Scholar] [CrossRef]
  49. Cullen, P.J.; Kaufman, C.K.; Bowman, W.C.; Kranz, R.G. Characterization of the Rhodobacter capsulatus Housekeeping RNA Polymerase. J. Biol. Chem. 1997, 272, 27266–27273. [Google Scholar] [CrossRef] [Green Version]
  50. Naylor, G.W.; Addlesee, H.A.; Gibson, L.C.D.; Hunter, C.N. The Photosynthesis Gene Cluster of Rhodobacter sphaeroides. Photosynth Res. 1999, 62, 121–139. [Google Scholar] [CrossRef]
  51. Alberti, M.; Burke, D.H.; Hearst, J.E. Structure and Sequence of the Photosynthesis Gene Cluster. In Anoxygenic Photosynthetic Bacteria; Kluwer Academic Publishers: Dordrecht, The Netherlands, 1995; pp. 1083–1106. [Google Scholar]
  52. Nakagawa, S.; Niimura, Y.; Miura, K.; Gojobori, T. Dynamic Evolution of Translation Initiation Mechanisms in Prokaryotes. Proc. Natl. Acad. Sci. USA 2010, 107, 6382–6387. [Google Scholar] [CrossRef]
  53. Friedman, D.I.; Pollara, B.; Gray, E.D. Structural Studies of the Ribosomes of Rhodopseudomonas spheroides. J. Mol. Biol. 1966, 22, 53–66. [Google Scholar] [CrossRef]
  54. Robinson, A.; Sykes, J. A Comparison of the Unfolding and Dissociation of the Large Ribosome Subunits from Rhodopseudomonas spheroides N.C.I.B. 8253 and Escherichia coli M.R.E. 600. Biochem. J. 1973, 133, 739–747. [Google Scholar] [CrossRef]
  55. Duval, M.; Korepanov, A.; Fuchsbauer, O.; Fechter, P.; Haller, A.; Fabbretti, A.; Choulier, L.; Micura, R.; Klaholz, B.P.; Romby, P.; et al. Escherichia coli Ribosomal Protein S1 Unfolds Structured mRNAs onto the Ribosome for Active Translation Initiation. PLoS Biol. 2013, 11, e1001731. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Brunel, C.; Romby, P.; Sacerdot, C.; De Smit, M.; Graffe, M.; Dondon, J.; Van Duin, J.; Ehresmann, B.; Ehresmann, C.; Springer, M. Stabilised Secondary Structure at a Ribosomal Binding Site Enhances Translational Repression in E. coli. J. Mol. Biol. 1995, 253, 277–290. [Google Scholar] [CrossRef] [PubMed]
  57. Emory, S.A.; Bouvet, P.; Belasco, J.G. A 5’-Terminal Stem-Loop Structure Can Stabilize mRNA in Escherichia coli. Genes Dev. 1992, 6, 135–148. [Google Scholar] [CrossRef] [Green Version]
  58. Arnold, T.E.; Yu, J.; Belasco, J.G. mRNA Stabilization by the OmpA 5’ Untranslated Region: Two Protective Elements Hinder Distinct Pathways for mRNA Degradation. RNA 1998, 4, 319–330. [Google Scholar] [PubMed]
  59. Chen, L.H.; Emory, S.A.; Bricker, A.L.; Bouvet, P.; Belasco, J.G. Structure and Function of a Bacterial mRNA Stabilizer: Analysis of the 5’ Untranslated Region of OmpA mRNA. J. Bacteriol. 1991, 173, 4578–4586. [Google Scholar] [CrossRef] [Green Version]
  60. Irastortza-Olaziregi, M.; Amster-Choder, O. Coupled Transcription-Translation in Prokaryotes: An Old Couple With New Surprises. Front. Microbiol. 2020, 11, 624830. [Google Scholar] [CrossRef]
  61. Petersen, C. Translation and mRNA Stability in Bacteria: A Complex Relationship. In Control of Messenger RNA Stability; Academic Press: Waltham, MA, USA, 1993; pp. 117–145. [Google Scholar]
  62. Iost, I.; Dreyfus, M. The Stability of Escherichia coli LacZ mRNA Depends upon the Simultaneity of Its Synthesis and Translation. EMBO J. 1995, 14, 3252–3261. [Google Scholar] [CrossRef]
Figure 1. Representative gene structure for translation enhancement. (a) Translation enhancement is observed when mlcR25 is inserted in the 5′-UTR [13]. (b) Nucleotide-sequence-containing mlcR25 (shown in magenta). The Shine–Dalgarno (SD) sequence is underlined. The start codon is shown in uppercase.
Figure 1. Representative gene structure for translation enhancement. (a) Translation enhancement is observed when mlcR25 is inserted in the 5′-UTR [13]. (b) Nucleotide-sequence-containing mlcR25 (shown in magenta). The Shine–Dalgarno (SD) sequence is underlined. The start codon is shown in uppercase.
Applmicrobiol 03 00047 g001
Figure 2. mlcR25 in the 5′-UTR prolongs mRNA longevity in E. coli. (a) Gene structures with or without mlcR25 under the control of lac promoter. (b) Relative amounts of indicated mRNAs over time after rifampicin treatment. The mean obtained from three independent experiments is shown. (c) Relative amounts of the indicated mRNAs obtained from the cells in the logarithmic phase. The mean obtained from three independent experiments is shown. p-values (unpaired t-test) are indicated.
Figure 2. mlcR25 in the 5′-UTR prolongs mRNA longevity in E. coli. (a) Gene structures with or without mlcR25 under the control of lac promoter. (b) Relative amounts of indicated mRNAs over time after rifampicin treatment. The mean obtained from three independent experiments is shown. (c) Relative amounts of the indicated mRNAs obtained from the cells in the logarithmic phase. The mean obtained from three independent experiments is shown. p-values (unpaired t-test) are indicated.
Applmicrobiol 03 00047 g002
Figure 3. TED is observed in the in vitro translation system. (a) Gene structures with or without mlcR25 under the control of the T7 promoter. (b) Western blot analysis of each reaction mixture using anti-GFP antibody. Plasmid DNA was added as the template. Arrowhead, GFP; asterisks, unknown proteins that were synthesized. (c) The relative band intensity of GFP normalized to bands of the control protein stained with CBB, related to (b). The mean obtained from three independent experiments is shown. (d) Western blot analysis of each reaction mixture using anti-GFP antibody. mRNA was added as the template. Arrowhead, GFP. (e) The relative band intensity of GFP normalized to bands of the control protein stained with CBB, related to (d). The mean obtained from three independent experiments is shown. p-values (one-way ANOVA followed by Tukey’s multiple comparisons test) are indicated.
Figure 3. TED is observed in the in vitro translation system. (a) Gene structures with or without mlcR25 under the control of the T7 promoter. (b) Western blot analysis of each reaction mixture using anti-GFP antibody. Plasmid DNA was added as the template. Arrowhead, GFP; asterisks, unknown proteins that were synthesized. (c) The relative band intensity of GFP normalized to bands of the control protein stained with CBB, related to (b). The mean obtained from three independent experiments is shown. (d) Western blot analysis of each reaction mixture using anti-GFP antibody. mRNA was added as the template. Arrowhead, GFP. (e) The relative band intensity of GFP normalized to bands of the control protein stained with CBB, related to (d). The mean obtained from three independent experiments is shown. p-values (one-way ANOVA followed by Tukey’s multiple comparisons test) are indicated.
Applmicrobiol 03 00047 g003
Figure 4. TED is applicable to the protein expression system of R. capsulatus. (a) Gene structures with or without mlcR25 under the control of sqr promoter, which is placed at 1238 bp upstream from the sqr start codon. (b) Nucleotide sequence containing mlcR25, which is shown in magenta. The SD sequence is underlined. The start codon is shown in uppercase. (c) Measurement of β-galactosidase activity driven by the sqr promoter. The enzymatic activity was calculated as the amount of o-nitrophenol (ONP) hydrolyzed from ONPG by β-galactosidase. The mean obtained from three independent experiments is shown. p-values (unpaired t-test) are indicated.
Figure 4. TED is applicable to the protein expression system of R. capsulatus. (a) Gene structures with or without mlcR25 under the control of sqr promoter, which is placed at 1238 bp upstream from the sqr start codon. (b) Nucleotide sequence containing mlcR25, which is shown in magenta. The SD sequence is underlined. The start codon is shown in uppercase. (c) Measurement of β-galactosidase activity driven by the sqr promoter. The enzymatic activity was calculated as the amount of o-nitrophenol (ONP) hydrolyzed from ONPG by β-galactosidase. The mean obtained from three independent experiments is shown. p-values (unpaired t-test) are indicated.
Applmicrobiol 03 00047 g004
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kondo, T.; Shimizu, T. Translation Enhancement by a Short Nucleotide Insertion at 5′UTR: Application to an In Vitro Cell-Free System and a Photosynthetic Bacterium. Appl. Microbiol. 2023, 3, 687-697. https://doi.org/10.3390/applmicrobiol3030047

AMA Style

Kondo T, Shimizu T. Translation Enhancement by a Short Nucleotide Insertion at 5′UTR: Application to an In Vitro Cell-Free System and a Photosynthetic Bacterium. Applied Microbiology. 2023; 3(3):687-697. https://doi.org/10.3390/applmicrobiol3030047

Chicago/Turabian Style

Kondo, Tomo, and Takayuki Shimizu. 2023. "Translation Enhancement by a Short Nucleotide Insertion at 5′UTR: Application to an In Vitro Cell-Free System and a Photosynthetic Bacterium" Applied Microbiology 3, no. 3: 687-697. https://doi.org/10.3390/applmicrobiol3030047

Article Metrics

Back to TopTop