Next Article in Journal
Anti-Migraine Effect of the Herbal Combination of Chuanxiong Rhizoma and Cyperi Rhizoma and UPLC-MS/MS Method for the Simultaneous Quantification of the Active Constituents in Rat Serum and Cerebral Cortex
Next Article in Special Issue
Molecular Characterization and Heterologous Production of the Bacteriocin Peocin, a DNA Starvation/Stationary Phase Protection Protein, from Paenibacillus ehimensis NPUST1
Previous Article in Journal
Comparison of Functional Components and Antioxidant Activity of Lycium barbarum L. Fruits from Different Regions in China
Previous Article in Special Issue
Efficient Conversion of Cane Molasses Towards High-Purity Isomaltulose and Cellular Lipid Using an Engineered Yarrowia lipolytica Strain in Fed-Batch Fermentation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Novel Cold-Adapted and Salt-Tolerant RNase R from Antarctic Sea-Ice Bacterium Psychrobacter sp. ANT206

Harbin Institute of Technology, School of Marine Science and Technology, Weihai 264209, China
*
Authors to whom correspondence should be addressed.
Molecules 2019, 24(12), 2229; https://doi.org/10.3390/molecules24122229
Submission received: 7 May 2019 / Revised: 10 June 2019 / Accepted: 11 June 2019 / Published: 14 June 2019
(This article belongs to the Special Issue Biotechnology of Microbial Enzymes)

Abstract

:
A novel RNase R, psrnr, was cloned from the Antarctic bacterium Psychrobacter sp. ANT206 and expressed in Escherichia coli (E. coli). A bioinformatics analysis of the psrnr gene revealed that it contained an open reading frame of 2313 bp and encoded a protein (PsRNR) of 770 amino acids. Homology modeling indicated that PsRNR had reduced hydrogen bonds and salt bridges, which might be the main reason for the catalytic efficiency at low temperatures. A site directed mutation exhibited that His 667 in the active site was absolutely crucial for the enzyme catalysis. The recombinant PsRNR (rPsRNR) showed maximum activity at 30 °C and had thermal instability, suggesting that rPsRNR was a cold-adapted enzyme. Interestingly, rPsRNR displayed remarkable salt tolerance, remaining stable at 0.5–3.0 M NaCl. Furthermore, rPsRNR had a higher kcat value, contributing to its efficient catalytic activity at a low temperature. Overall, cold-adapted RNase R in this study was an excellent candidate for antimicrobial treatment.

1. Introduction

Ribonuclease (Ribonuclease, RNase), an important nucleic acid hydrolase in organism, can hydrolyze the phosphodiester bonds among the nucleic acid residues of RNA and play a significant role in different stages of the cell metabolism [1,2]. In Escherichia coli (E. coli), eight distinct RNases involved in different aspects of RNA metabolism were identified. RNase R, originally termed as vacB in Shigella flexneri [3], was found in Aeromonas hydrophila [4], Legionella pneumophila [5], and Pneumoniae [6]. RNase R participates in an essential cell function besides its role in virulence [3]. It has been reported that RNase R can not only participate in controlling the type and quantity distribution of intracellular RNA but degrade different types of RNA [7]. RNase R’s interaction with ribosomal proteins might be a result of this enzyme’s involvement in the ribosome quality control [8,9]. A growing number of evidence has indicated that RNase R is a cold shock protein (CspA), and its expression regulation is closely related to microbial adaptation [10,11]. It is well-known that the recruitment of RNase R to the degradosome complex is related to its need for low temperatures during growth [12]. It has been discovered that RNase R in Pseudomonas syringae plays an important role in growth at low temperature [13]. More importantly, RNase R is necessary for maintaining the precise amount of active ribosomal complexes required for proper mRNA translation, and, accordingly, RNase R might be a promising target for antimicrobial treatment [4]. Due to the function of RNase R in RNA metabolism and biological activities [3,12], it might play an important role in the understanding of virulence associated with enterobacteria and have great value in antimicrobial treatment.
The unique geographical and climatic characteristics of Antarctic sea-ice have produced a wealth of extreme microbial resources. In the long-term evolution, microbes gave birth to very unique genetic resources, genetic backgrounds, and metabolic characteristics. Recently, it has been demonstrated that RNA transcription factors and protein synthetic factors play a vital part in Antarctic bacterium low temperature adaptation mechanisms [14]. Due to it being key enzyme in the regulation of RNA metabolism system, a study concerning the structural and biological characteristics of RNase R from Antarctic organisms is imminent. The present study was aimed at conducting molecular and enzymatic characterizations of the cold-adapted RNase R to have an insight into its structural and catalytic properties.

2. Results and Discussion

2.1. Identification of psrnr Gene

The psrnr gene was seen to be 2313 bp in length (GenBank accession numbers MK624989) and to encode a protein with 770 amino acids, while the length of the rnr gene from Pseudomonas syringae Lz4W was seen to be 2658 bp, encoding 885 amino acids [12]. Based on sequence alignments with the related RNase Rs, the residues (Thr 646, Phe 654, Leu 665 and His 667) were identified as the catalytic sites in the sequence of PsRNR (Figure 1). However, RNase R from E. coli DEC6A was seen to have two catalytic sites (Asp 272 and Asp 281) [15]. Besides, it was reported that two catalytic sites from E. coli were identified in Tyr 324 and Asp 280 [16]. Another RNase R from E. coli contained three catalytic sites (Asp 272, Asp 278, and Asp 280) [17]. In addition, the sequence homology showed that PsRNR displayed the highest similarity (90.07%) to RNase R from Psychrobacter cryohalolentis, followed by the RNase R from Psychrobacter arcticus (88.98%).

2.2. Homology 3D Modeling and Structure Analysis

For the validation of the structural model, 95.89% of the residues had an averaged 3D-1D score ≥ 0.2. The model was validated by a Ramachandran plot analysis with 90.8% residues in the most favored regions. These parameters indicated that the structural model of PsRNR was well qualified. As shown in Figure 2, the structural model of PsRNR superimposed well with the homologous RNase R (EcRNase R, PDB: 5XGU). Furthermore, in comparing the structural features between PsRNR and EcRNase R, it can be seen that the former showed several cold-adapted characteristics (Table 1). Firstly, in comparison with its mesophilic homologues, PsRNR showed less electrostatic interactions, particularly hydrogen bonds and salt bridges, resulting in a decrease in stability and thermostability [18,19]. A similar study on cold-adapted chitinase also revealed that the reduction of electrostatic interactions contributed to its thermal instability [20]. Secondly, less hydrophobic interactions might also be advantageous to the structural flexibility of cold-adapted enzymes [21], and PsRNR possessed reduced hydrophobic interactions (447) in comparison with EcRNase R (543). Furthermore, PsRNR showed a lower Arg/(Arg + Lys) ratio, which manifested as increased conformational flexibility and low temperature catalytic competence [22]. Similar structural features were also obtained in the study of the cold-adapted esterase [23].

2.3. Expression, Purification and Enzyme Assays

In comparing BL21/pET-28a(+) with isopropyl β-D-thiogalactoside (IPTG) induction (Figure 3, Lane 2), one can see that the crude extract from the recombinant bacteria displayed a different band with an approximately molecular mass of 91.4 kDa (Figure 3, Lane 3, red arrow), indicating that PsRNR was expressed in E. coli BL21. The purified rPsRNR displayed a major band on SDS-PAGE (Figure 3, Lanes 4–6). The rPsRNR was purified approximately 5.58-fold with a recovery yield of 48.85%, and its specific activity was 115.60 μmol/min/mg; that of RNase R from E. coli was 13-fold with a recovery yield of 40% [7].

2.4. Site-Directed Mutagenesis

The activity of the mutant rPsRNR generated by site-directed mutagenesis was investigated. As shown in Figure 4, the loss of the enzymatic activity of three mutants T646A, F654A, and L665A, was 12%, 55%, and 50%, respectively, confirming that these residues might participate in the process of RNase R catalysis [24]. Importantly, the H667A mutant protein completely lost the activity toward yeast RNA, indicating that His667 was necessary for catalysis. Nevertheless, residues of D272 and D280 in E. coli RNase R were crucial for its catalytic activity [17], and the residue of D280 in RNase R played a key role in its RNase activity [16].

2.5. Biochemical Characteristics of rPsRNR

The optimal temperature of rPsRNR was 30 °C (Figure 5a), while the RNase R from the Antarctic bacterium Pseudomonas syringae showed maximal activity at 25 °C [12]. Furthermore, the maximal activity of an extracellular RNase from Bacillus cereus was at 60 °C [25]. Importantly, rPsRNR, which had lower activity than psychrophilic RNase R, retained 22.0–38.0% of its initial activity even at 0–10 °C [12]. For the enzymatic thermostability, the half-life of the rPsRNR activity was approximately 22 min at 50 °C, but rPsRNR was completely inactivated for 70 min at 50 °C (Figure 5b). In contrast, RNase R from E. coli lost approximately 80% of its full activity when incubated 5 min at 50 °C [12]. RNase II, an exoribonuclease homologous to RNase R, was most active at 50 °C for 5 min [7]. The optimum pH for rPsRNR activity was 6.0 (Figure 5c), while the iso-Ribonucleases from thermophilic fungus showed full activity at pH 3.0 [26]. In addition, RNase II from E. coli displayed a broad optimal pH range between 7.5 and 9.5 [7], and the activity of RNase R from Mycoplasma genitalium was maximum at pH 8.5; the activity dropped by about 20% at pH 7.5 and 9.0 [27]. As for the pH stability, rPsRNR maintained about 80.0% of its maximal activity over the pH range of 5.5 to 6.5 at 30 °C. (Figure 5d), suggesting that rPsRNR exhibited good stability under the weak acid condition. On the contrary, RNase HII from Aeropyrum pernix exhibited maximal activity at an alkaline pH [28]. This range of pH dependence for activity and stability means that rPsRNR could potentially contribute to the medical industry. The activity of rPsRNR was investigated in the presence of various concentrations of NaCl (Figure 5e). rPsRNR kept its highest activity in the presence of 1.5 M NaCl and kept over 85.0% of its activity at 0.5–3.0 M NaCl; it lost approximately 36.9% of its activity at the concentration of 4.0 M. This data reflects that rPsRNR features salt tolerance characteristics, which may be related to the high salt concentration in Antarctic sea ice. Based on rPsRNR’s activity on the effects of various compounds (Table 2), rPsRNR was strongly inhibited by 1 mM Ba2+, and the rPsRNR activity was not detected after being incubated in 5 mM Ba2+ for 30 min. In addition, the inhibitions by 1 mM Cu2+ and Cr2+ were 59.1% and 36.8%, respectively. It was noted that activity was enhanced by 21.4% and 31.2% after adding 1 mM and 5 mM Mg2+, respectively, and, thus, it was reasonable to assume that rPsRNR required a certain concentration of Mg2+ to improve its activity. Similarly, E. coli and Mycoplasma genitalium RNase R were found to be most active at 0.1–0.5 mM Mg2+ [7,27]. Besides, the activity of RNase H3 from Aquifex aeolicus was improved in the presence of Mg2+ [29]. Furthermore, 1 mM of ethylenediamine tetraacetic acid (EDTA) inhibited rPsRNR activity completely, which was similar to the report of RNase R from E. coli [7] and Mycoplasma genitalium [27].

2.6. Enzyme Kinetics of rPsRNR

Kinetics parameters (Km, Vm, and kcat) of rPsRNR were measured at different temperatures (Figure 6). As shown in Table 3, the Km values of rPsRNR decreased remarkably from 1.008 to 0.339 μM from 0 to 30 °C. The Km trend of endonuclease I was the same as rPsRNR [30]. However, the Km values of cold-adapted invertase increased with increasing temperatures, and Km values were higher than its thermostable counterparts to decrease the activation free-energy barrier [31]. The Km and Vmax values of RNase HII from Aeropyrum pernix were 0.361 mM and 0.196 mM/min at 30 °C, respectively [28]. In addition, an extracellular ribonuclease from a Bacillus sp. exhibited a Km value of 0.12 mg/mL and a Vmax value of 55.5 μg/mL/min at 37 °C [32]. In addition, the kcat values at 0 °C and 30 °C were 27.126 s−1 and 70.767 s−1, respectively. In generally, cold-adapted enzymes were quite efficient in compensating for the reduced reaction rates at low temperatures by improving of the kcat value [33]. Therefore, it is speculated that rPsRNR, like endonuclease I from cold-adapted Vibrio salmonicida [30], might adapt to low temperatures by increasing kcat. Obviously, as the temperature increased, the kcat gradually increased in a similar fashion to the tendency of cold-adapted β-d-galactosidase to kcat at 10–30 °C [34]. Overall, the high catalytic efficiency at low temperatures of rPsRNR may make it a potential candidate for molecular biology applications.

2.7. Enzyme Thermodynamics of rPsRNR

Thermodynamic parameters (∆H, ∆S, and ∆G) of the rPsRNR were also measured (Table 4). The ∆G values of the rPsRNR increased from 59.20 to 63.55 kJ/mol when the temperature increased from 0 to 30 °C. This trend has been found in the xylanase rXynAGN16L from Arthrobacter sp. GN16 [35]. However, with the increase of temperature from 45 to 60 °C, the ∆G values of cold-adapted endoglucanase exhibited a downward trend, indicating that the thermostability of endoglucanase decreased with increasing temperature [36]. Furthermore, the ∆H values of the rPsRNR decreased from 0 to 30 °C. In general, the decrease of ∆H weakened the exponential temperature dependence of the reaction rate, thus contributing to the catalysis at low temperature. Therefore, rPsRNR showed smaller ∆H than psychrophilic endonuclease I and cold-adapted xylanase rXynAGN16L [30,35].

3. Materials and Methods

3.1. Strains and Bacteria Cultivation

The RNase R-producing strain Psychrobacter sp. ANT206 (GenBank accession numbers MK968312) was isolated from Antarctic sea-ice (68° 300 E, 65° 000 S). Strain ANT206 was cultured in a 2216E medium at 12 °C and was shook at 200 rpm. pET-28(a) was stored in our lab and used as the vector for protein expression. E. coli BL21 was used as the host for protein expression. E. coli strains and recombinant bactera were cultured in a Luria Bertani (LB) medium containing kanamycin (100 mg/L).

3.2. Identification of psrnr Gene

According to the sequence and annotation of the genome of Psychrobacter sp. ANT206 (data not shown), the forward primers 5′-ACTGGATCCATGT CAAACCAAGATC-3′ (BamHI site underlined) and the reverse primers 5′-TACCTCGAGCGCTCTTTTTACTACT-3′ (XhoI site underlined) were designed and amplified by PCR for the full-length RNase R gene (psrnr). The complete amino acid sequences of the PsRNR were obtained through an open reading frame finder (ORF finder). The catalytic sites were predicted by CD-Search (https://www.ncbi.nlm.nih.gov/Structure/cdd/wrpsb.cgi). Multiple sequence alignments were performed using the Bioedit and ESPript program (http://espript.ibcp.fr/ESPript/cgi-bin/ESPript.cgi).

3.3. Homology Modeling of PsRNR

The 3D structure of the PsRNR was established by the SWISS-MODEL server and verified using SAVES V 5.0 (http://servicesn.mbi.ucla.edu/SAVES/) (University of California, Los Angeles, CA, USA). PyMOL 2.2.0 software (DeLano Scientific LLC, San Carlos, CA, USA) was used to visualize homology modeling with EcRNase R (PDB ID: 5XGU) as a template. In addition, salt bridges were predicted to use the VMD1.9.3 (University of lllinois, Urbana–Champaign, USA). Meanwhile, protein intramolecular interactions (hydrogen bonds, cation-pi interactions, hydrophobic interactions, and aromatic interactions) were predicted by the Protein Interactions Calculator program (http://pic.mbu.iisc.ernet.in/job.html).

3.4. Site-Directed Mutagenesis

Using the vector pET-28a (+) as a temple, the site-directed mutagenesis of T646, F654, L665, and H667 to Ala was performed with the QuikChange site-directed mutagenesis kit (Stratagene) by following the manufacturer’s protocol. The mutant primers were designed to amplify the DNA sequence, and the mutagenesis sites were underlined (Table 5).

3.5. Expression and Purification of PsRNR

In order to clone the psrnr gene and its mutant genes into the pET-28a(+) expression vector, the PCR product was inserted directly into pET-28a(+) vector. The recombinant plasmid was transformed into receptor E. coli BL21, and the sequences were then verified by sequencing. The recombinant strains grew in a LB medium and cultured at 37 °C until the OD600 nm was 0.6–0.8. After the optimization of protein expression, 238 mg/L IPTG was added for induction at 20 °C for 16 h. The induced cells were centrifuged and disrupted by ultrasonication (JY96-IIN, Shanghai, China). Then, the inclusion bodies were washed for removing the insoluble debris, treated with 8 M urea, and centrifuged at 7500× g for 15 min. The supernatant was diluted by 30 times by adding Tris-HCl (pH 8.0) at 25 °C for 2 h, which was the procedure of protein refolding. The protein solution was centrifuged at 12,000× g for 15 min, and the supernatant was the crude extract of proteins. Purification of rPsRNR and its mutant proteins were employed by Ni-NTA affinity chromatography (GE Healthcare, Uppsala, Sweden). The samples were eluted with 50 mM imidazole buffer (20 mM Tris-HCl, 500 mM NaCl, pH 8.0) at a flow rate of 0.5 mL/min. The purity and molecular mass of rPsRNR were determined by 12.0% polyacrylamide gels SDS-PAGE.

3.6. Enzyme Assay

The activity of rPsRNR toward yeast RNA was determined based on the previous method [37]. The reaction system contained a 0.1 M NaAc-HAc buffer (pH 6.0), 1.25 μM of yeast RNA, and 7.5 μg of purified protein. The reaction was performed for 10 min at specified temperatures and stopped by adding 350 μL ice–cold 3.4% perchloric acid. One unit of enzyme activity was defined as the amount of enzyme that brought about an increase in absorbance at 260 nm of 1 unit (U) per min in a cuvette of 1 cm path length of per mg of reaction system under the specified conditions.

3.7. Biochemical Characteristics of rPsRNR

The purified protein (7.2 μg) was added in the reaction system to investigate its biochemical characteristics. To evaluate the optimal temperature of the rPsRNR, enzymatic activity was carried out at various temperatures (0–50 °C). For the enzymatic thermostability, purified rPsRNR was incubated at 40, 45, and 50 °C for 90 min. The optimum pH of rPsRNR was determined at 30 °C in NaAc/HAc (pH 4.0–6.0) and Na2HPO4/NaH2PO4 (pH 6.0–8.5) buffers. To test pH stability, purified rPsRNR was incubated in different pH buffers (4.0–8.5) at 30 °C for 30 min. To investigate the salt tolerance, purified rPsRNR was incubated at 30 °C for 30 min in 0–4.0 M NaCl buffers, and then the residual activity was assayed. The effects of various compounds on rPsRNR activity were investigated by the purified rPsRNR in buffer containing different reagents at 30 °C for 30 min, and the residual activity was assayed.

3.8. Kinetic Parameters and Thermodynamic Parameters of the rPsRNR

The reaction system of kinetic parameters contained 0.1 M of a NaAc–HAc buffer (pH 6.0), different concentrations of yeast RNA (0.25, 0.50, 0.75, 1.00, 1.25, and 1.5 μM), and purified rPsRNR (6.8 μg), all of which incubated at 0–30 °C for 10 min. Then, the reaction was stopped by adding 350 μL of ice-cold 3.4% perchloric acid. Km and Vmax were measured using the Michaelis–Menten equation [38]. The kcat was calculated by the determination of kinetics, and the thermodynamic parameters (∆H, ∆S, and ∆G) were determined by the modification method [38].

4. Conclusions

A novel RNase R gene (psrnr) was cloned, expressed, and characterized in this study. PsRNR exhibited the structural characteristics of cold-adapted enzymes by homology modeling, including reduced hydrogen bonds and salt bridges. Furthermore, the residue of His667 was found to be necessary for the catalysis of PsRNR. After purification, rPsRNR exhibited maximal activity at 30 °C and pH 6.0, and it had excellent salt tolerance. Additionally, the thermodynamic characterization reflected the lower values of ∆H, ∆S, and ∆G at lower temperatures. Given these interesting characteristics, rPsRNR would be a potential candidate for antimicrobial treatment.

Author Contributions

Methodology, Y.W. (Yatong Wang), Y.H., and Q.W. (Quanfu Wang); Data curation Y.W. (Yatong Wang), Y.H., P.N., Y.W. (Yifan Wang), X.R., and Q.W. (Qifeng Wei); writing-original draft preparation, Y.W. (Yatong Wang) and Q.W (Quanfu Wang).

Funding

This research was funded by the National Science Foundation of China (Grant No. 41876149), the Natural Science Foundation of Shandong Province (Grant No. ZR2017MC046 and ZR2019MD018) and the Key Research and Development Plan of Shandong Province (Grant No. 2018GHY115021).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fong, W.P.; Mock, W.Y.; Ng, T.B. Intrinsic ribonuclease activities in ribonuclease and ribosome-inactivating proteins from the seeds of bitter gourd. Int. J. Biochem. Cell Biol. 2000, 32, 571–577. [Google Scholar] [CrossRef]
  2. Cheng, Z.F.; Deutscher, M.P. An important role for RNase R in mRNA decay. Mol. Cell 2005, 17, 313–318. [Google Scholar] [CrossRef] [PubMed]
  3. Cheng, Z.F.; Zuo, Y.H.; Li, Z.W.; Rudd, K.E.; Deutscher, M.P. The vacB gene required for virulence in Shigella flexneri and Escherichia coli encodes the exoribonuclease RNase R. J. Biol. Chem. 1998, 273, 14077–14080. [Google Scholar] [CrossRef] [PubMed]
  4. Erova, T.E.; Kosykh, V.G.; Fadl, A.A.; Sha, J.; Horneman, A.J.; Chopra, A.K. Cold shock exoribonuclease R (vacB) is involved in Aeromonas hydrophila pathogenesis. J. Bacteriol. 2008, 190, 3467–3474. [Google Scholar] [CrossRef] [PubMed]
  5. Charpentier, X.; Faucher, S.P.; Kalachikov, S.; Shuman, H.A. Loss of RNase R induces competence development in Legionella pneumophila. J. Bacteriol. 2008, 190, 8126–8136. [Google Scholar] [CrossRef] [PubMed]
  6. Barria, C.; Domingues, S.; Arraiano, C.M. Pneumococcal RNase R globally impacts protein synthesis by regulating the amount of actively translating ribosomes. RNA Biol. 2019, 16, 211–219. [Google Scholar] [CrossRef] [PubMed]
  7. Cheng, Z.F.; Deutscher, M.P. Purification and characterization of the Escherichia coli exoribonuclease RNase R—Comparison with RNase II. J. Biol. Chem. 2002, 277, 21624–21629. [Google Scholar] [CrossRef] [PubMed]
  8. Malecki, M.; Barria, C.; Arraiano, C.M. Characterization of the RNase R association with ribosomes. BMC Microbiol. 2014, 14. [Google Scholar] [CrossRef]
  9. Domingues, S.; Moreira, R.N.; Andrade, J.M.; dos Santos, R.F.; Barria, C.; Viegas, S.C.; Arraiano, C.M. The role of RNase R in trans-translation and ribosomal quality control. Biochimie 2015, 114, 113–118. [Google Scholar] [CrossRef]
  10. Phadtare, S. Escherichia coli cold-shock gene profiles in response to over-expression/deletion of CsdA, RNase R and PNPase and relevance to low-temperature RNA metabolism. Genes Cells 2012, 17, 850–874. [Google Scholar] [CrossRef]
  11. Moreira, R.N.; Domingues, S.; Viegas, S.C.; Amblar, M.; Arraiano, C.M. Synergies between RNA degradation and trans-translation in Streptococcus pneumoniae: Cross regulation and co-transcription of RNase R and SmpB. BMC Microbiol. 2012, 12. [Google Scholar] [CrossRef]
  12. Sulthana, S.; Rajyaguru, P.I.; Mittal, P.; Ray, M.K. rnr Gene from the Antarctic Bacterium Pseudomonas syringae Lz4W, Encoding a Psychrophilic RNase R. Appl. Environ. Microbiol. 2011, 77, 7896–7904. [Google Scholar] [CrossRef]
  13. Purusharth, R.I.; Madhuri, B.; Ray, M.K. Exoribonuclease R in Pseudomonas syringae is essential for growth at low temperature and plays a novel role in the 3′ end processing of 16 and 5S ribosomal RNA. J. Biol. Chem. 2007, 282, 16267–16277. [Google Scholar] [CrossRef]
  14. Piette, F.; D’Amico, S.; Struvay, C.; Mazzucchelli, G.; Renaut, J.; Tutino, M.L.; Danchin, A.; Leprince, P.; Feller, G. Proteomics of life at low temperatures: Trigger factor is the primary chaperone in the Antarctic bacterium Pseudoalteromonas haloplanktis TAC125. Mol. Microbiol. 2010, 76, 120–132. [Google Scholar] [CrossRef]
  15. Chu, L.Y.; Hsieh, T.J.; Golzarroshan, B.; Chen, Y.P.; Agrawal, S.; Yuan, H.N.S. Structural insights into RNA unwinding and degradation by RNase R. Nucleic Acids Res. 2017, 45, 12015–12024. [Google Scholar] [CrossRef] [Green Version]
  16. Matos, R.G.; Barbas, A.; Arraiano, C.M. RNase R mutants elucidate the catalysis of structured RNA: RNA-binding domains select the RNAs targeted for degradation. Biochem. J. 2009, 423, 291–301. [Google Scholar] [CrossRef]
  17. Awano, N.; Rajagopal, V.; Arbing, M.; Patel, S.; Hunt, J.; Inouye, M.; Phadtare, S. Escherichia coli RNase R Has Dual Activities, Helicase and RNase. J. Bacteriol. 2010, 192, 1344–1352. [Google Scholar] [CrossRef]
  18. Zheng, Y.Y.; Li, Y.J.; Liu, W.D.; Chen, C.C.; Ko, T.P.; He, M.; Xu, Z.X.; Liu, M.X.; Luo, H.Y.; Guo, R.T.; et al. Structural insight into potential cold adaptation mechanism through a psychrophilic glycoside hydrolase family 10 endo-beta-1,4-xylanase. J. Struct. Biol. 2016, 193, 206–211. [Google Scholar] [CrossRef]
  19. Kumar, S.; Tsai, C.J.; Nussinov, R. Factors enhancing protein thermostability. Protein Eng. 2000, 13, 179–191. [Google Scholar] [CrossRef] [Green Version]
  20. Ramli, A.N.M.; Mahadi, N.M.; Shamsir, M.S.; Rabu, A.; Joyce-Tan, K.H.; Murad, A.M.A.; Illias, R.M. Structural prediction of a novel chitinase from the psychrophilic Glaciozyma antarctica PI12 and an analysis of its structural properties and function. J. Comput. Aided Mol. Des. 2012, 26, 947–961. [Google Scholar] [CrossRef]
  21. Siddiqui, K.S.; Cavicchioli, R. Cold-adapted enzymes. Annu. Rev. Biochem. 2006, 75, 403–433. [Google Scholar] [CrossRef]
  22. Feller, G.; Gerday, C. Psychrophilic enzymes: Hot topics in cold adaptation. Nat. Rev. Microbiol. 2003, 1, 200–208. [Google Scholar] [CrossRef]
  23. Fu, J.; Leiros, H.K.S.; de Pascale, D.; Johnson, K.A.; Blencke, H.M.; Landfald, B. Functional and structural studies of a novel cold-adapted esterase from an Arctic intertidal metagenomic library. Appl. Microbiol. Biotechnol. 2013, 97, 3965–3978. [Google Scholar] [CrossRef]
  24. Hossain, S.T.; Malhotra, A.; Deutscher, M.P. How RNase R Degrades Structured RNA: Role of the helicase activity and the S1 domain. J. Biol. Chem. 2016, 291, 7877–7887. [Google Scholar] [CrossRef]
  25. Zhou, W.W.; Niu, T.G. Purification and some properties of an extracellular ribonuclease with antiviral activity against tobacco mosaic virus from Bacillus cereus. Biotechnol. Lett. 2009, 31, 101–105. [Google Scholar] [CrossRef]
  26. Landry, K.S.; Levin, R.E. Purification and Characterization of Iso-Ribonucleases from a Novel Thermophilic Fungus. Int. J. Mol. Sci. 2014, 15, 944–957. [Google Scholar] [CrossRef] [Green Version]
  27. Lalonde, M.S.; Zuo, Y.; Zhang, J.; Gong, X.; Wu, S.; Malhotra, A.; Li, Z. Exoribonuclease R in Mycoplasma genitalium can carry out both RNA processing and degradative functions and is sensitive to RNA ribose methylation. RNA 2007, 13, 1957–1968. [Google Scholar] [CrossRef]
  28. Hou, J.L.; Liu, Y.F.; Lu, Z.; Liu, X.P.; Liu, J.H. Biochemical characterization of RNase HII from Aeropyrum pernix. Acta Biochim. Biophys. Sin. 2012, 44, 339–346. [Google Scholar] [CrossRef]
  29. Jongruja, N.; You, D.J.; Angkawidjaja, C.; Kanaya, E.; Koga, Y.; Kanaya, S. Structure and characterization of RNase H3 from Aquifex aeolicus. FEBS J. 2012, 15, 2737–2753. [Google Scholar] [CrossRef]
  30. Altermark, B.; Niiranen, L.; Willassen, N.P.; Smalas, A.O.; Moe, E. Comparative studies of endonuclease I from cold-adapted Vibrio salmonicida and mesophilic Vibrio cholerae. FEBS J. 2007, 274, 252–263. [Google Scholar] [CrossRef]
  31. Zhou, J.P.; He, L.M.; Gao, Y.J.; Han, N.Y.; Zhang, R.; Wu, Q.; Li, J.J.; Tang, X.H.; Xu, B.; Ding, J.M.; et al. Characterization of a novel low-temperature-active, alkaline and sucrose-tolerant invertase. Sci. Rep. 2016, 6. [Google Scholar] [CrossRef]
  32. Shruti, G.; Sukhdev, S.; Singh, K.S. Purification and characterization of an extracellular ribonuclease from a Bacillus sp. RNS3 (KX966412). Int. J. Biol. Macromol. 2017, 97, 440–446. [Google Scholar] [CrossRef]
  33. Hoyoux, A.; Jennes, I.; Dubois, P.; Genicot, S.; Dubail, F.; Francois, J.M.; Baise, E.; Feller, G.; Gerday, C. Cold-adapted beta-galactosidase from the Antarctic psychrophile Pseudoalteromonas haloplanktis. Appl. Environ. Microbiol. 2001, 67, 1529–1535. [Google Scholar] [CrossRef]
  34. Pawlak-Szukalska, A.; Wanarska, M.; Popinigis, A.T.; Kur, J. A novel cold-active β-D-galactosidase with transglycosylation activity from the Antarctic Arthrobacter sp. 32cB-Gene cloning, purification and characterization. Process Biochem. 2014, 49, 2122–2133. [Google Scholar] [CrossRef]
  35. Zhou, J.P.; Liu, Y.; Shen, J.D.; Zhang, R.; Tang, X.H.; Li, J.J.; Wang, Y.Y.; Huang, Z. X Kinetic and thermodynamic characterization of a novel low-temperature-active xylanase from Arthrobacter sp. GN16 isolated from the feces of Grus nigricollis. Bioengineered 2015, 6, 111–114. [Google Scholar] [CrossRef]
  36. Xue, D.S.; Zeng, X.H.; Gong, C.J.; Lin, D.Q.; Yao, S.J. A cold adapt and ethanol tolerant endoglucanase from a marine Bacillus subtilis. Chin. J. Chem. Eng. 2018, 26, 2601–2606. [Google Scholar] [CrossRef]
  37. Kunitz, M. A spectrophotometric method for the measurement of ribonuclease activity. J. Biol. Chem. 1946, 164, 563–568. [Google Scholar]
  38. Lonhienne, T.; Gerday, C.; Feller, G. Psychrophilic enzymes: Revisiting the thermodynamic parameters of activation may explain local flexibility. Biochim. Biophys. Acta 2000, 1543, 1–10. [Google Scholar] [CrossRef]
Sample Availability: Samples of the recombinant strain is available from the authors.
Figure 1. Alignment of the PsRNR sequence with other RNase R. The species names and GenBank accession numbers are as follows: PsRNR, Psychrobacter sp. ANT206 RNase R (MK624989), EcRNase R, Escherichia coli RNase R (WP_038432731, PDB ID: 5XGU), PaRNase R, Psychrobacter arcticus RNase R (WP_011281354), PgRNase R, Psychrobacter glacincola RNase R (WP_055125677), MoRNase R, Moraxella osloensis RNase R (WP_065264345), and MlRNase R, Moraxella lincolnii RNase R (WP_078306359). The conserved catalytic sites are indicated with cycles.
Figure 1. Alignment of the PsRNR sequence with other RNase R. The species names and GenBank accession numbers are as follows: PsRNR, Psychrobacter sp. ANT206 RNase R (MK624989), EcRNase R, Escherichia coli RNase R (WP_038432731, PDB ID: 5XGU), PaRNase R, Psychrobacter arcticus RNase R (WP_011281354), PgRNase R, Psychrobacter glacincola RNase R (WP_055125677), MoRNase R, Moraxella osloensis RNase R (WP_065264345), and MlRNase R, Moraxella lincolnii RNase R (WP_078306359). The conserved catalytic sites are indicated with cycles.
Molecules 24 02229 g001
Figure 2. The 3D structure models of PsRNR and structure superimposition with EcRNase R (PDB ID: 5XGU). The catalytic triad residues are indicated as stick models.
Figure 2. The 3D structure models of PsRNR and structure superimposition with EcRNase R (PDB ID: 5XGU). The catalytic triad residues are indicated as stick models.
Molecules 24 02229 g002
Figure 3. SDS-PAGE analysis of expression and purification of the rPsRNR. Lane 1, protein molecular weight marker; Lane 2, crude extract from the BL21/pET-28a(+)with IPTG induction; Lane 3, crude extract from the BL21/pET-28a(+)-rPsRNR; Lane 4–6, the purified rPsRNR protein by Ni-NTA using the buffer (pH 8.0) contained 20 mM Tris-HCl, 100 mM NaCl and 50 mM imidazole.
Figure 3. SDS-PAGE analysis of expression and purification of the rPsRNR. Lane 1, protein molecular weight marker; Lane 2, crude extract from the BL21/pET-28a(+)with IPTG induction; Lane 3, crude extract from the BL21/pET-28a(+)-rPsRNR; Lane 4–6, the purified rPsRNR protein by Ni-NTA using the buffer (pH 8.0) contained 20 mM Tris-HCl, 100 mM NaCl and 50 mM imidazole.
Molecules 24 02229 g003
Figure 4. Measurement of the relative activity of the mutant rPsRNR generated by site-directed mutagenesis. The activity of the enzyme without mutation was defined as 100%.
Figure 4. Measurement of the relative activity of the mutant rPsRNR generated by site-directed mutagenesis. The activity of the enzyme without mutation was defined as 100%.
Molecules 24 02229 g004
Figure 5. Biochemical characteristics of rPsRNR. (a) The optimal temperature was determined by measuring the activity at temperatures from 0 to 50 °C. Its maximal activity was taken as 100%. (b) Effect of temperatures on the stability of the purified rPsRNR. The enzyme was incubated at 40 °C (■), 45 °C (●), and 50 °C (▲), for 90 min. Its maximal activity was taken as 100%. (c) The optimal pH was determined by measuring the activity at pH from 4.0 to 8.5. The reaction contained NaAc/HAc(■) and Na2HPO4/ Na2HPO4 (●). (d) Effect of pH on the stability of the purified RNase R. The enzyme was incubated at 30 °C for 30 min. Its maximal activity was taken as 100%. (e) The enzyme was incubated by different concentrations of NaCl for 30 min. Its activity with 0 M NaCl was set as 100%.
Figure 5. Biochemical characteristics of rPsRNR. (a) The optimal temperature was determined by measuring the activity at temperatures from 0 to 50 °C. Its maximal activity was taken as 100%. (b) Effect of temperatures on the stability of the purified rPsRNR. The enzyme was incubated at 40 °C (■), 45 °C (●), and 50 °C (▲), for 90 min. Its maximal activity was taken as 100%. (c) The optimal pH was determined by measuring the activity at pH from 4.0 to 8.5. The reaction contained NaAc/HAc(■) and Na2HPO4/ Na2HPO4 (●). (d) Effect of pH on the stability of the purified RNase R. The enzyme was incubated at 30 °C for 30 min. Its maximal activity was taken as 100%. (e) The enzyme was incubated by different concentrations of NaCl for 30 min. Its activity with 0 M NaCl was set as 100%.
Molecules 24 02229 g005
Figure 6. Lineweaver–Burk double reciprocal plots of purified RNase R with respect to yeast RNA at 0 °C (■), 5 °C (○), 10 °C (▲), 15 °C (□), 20 °C (●), 25 °C (△), and 30 °C (◆). Activity was measured as above described. Every experiment was done in triplicates.
Figure 6. Lineweaver–Burk double reciprocal plots of purified RNase R with respect to yeast RNA at 0 °C (■), 5 °C (○), 10 °C (▲), 15 °C (□), 20 °C (●), 25 °C (△), and 30 °C (◆). Activity was measured as above described. Every experiment was done in triplicates.
Molecules 24 02229 g006
Table 1. Comparison of structural adaption features between PsRNR and its homolog (EcRNase R).
Table 1. Comparison of structural adaption features between PsRNR and its homolog (EcRNase R).
ParametersPsRNREcRNase RExpected Effect on PsRNR
Salt Bridges1727Stability
Hydrogen Bonds478492
Cation-pi interactions1114
Aromatic 2223
Hydrophobic interactions 447543Thermolability
Glycine residues4659Flexibility
Proline residues3735
Arginine residues4474
Arg/(Arg + Lys)0.470.56
Table 2. Effect of various reagents on the rPsRNR activity.
Table 2. Effect of various reagents on the rPsRNR activity.
ReagentConcentrationRelative Activity (%)ReagentConcentrationRelative Activity (%)
None--100.0None--100.0
Mg2+1 mM121.4 ± 2.6Mg2+5 mM131.2 ±4.8
Ca2+1 mM102.3 ± 4.3Ca2+5 mM86.0 ± 3.5
Zn2+1 mM81.8 ± 2.3Zn2+5 mM75.8 ± 1.5
Fe2+1 mM90.9 ± 6.2Fe2+5 mM85.9 ± 4.1
Cu2+1 mM40.9 ± 1.9Cu2+5 mM20.5 ± 2.5
Pb2+1 mM72.7 ± 5.6Pb2+5 mM65.3 ± 3.7
Cr2+1 mM63.2 ± 3.7Cr2+5 mM62.7 ± 5.5
Ba2+1 mM2.1 ± 2.9Ba2+5 mMND
EDTA1 mMNDEDTA5 mMND
ND: Activity was not detected.
Table 3. The kinetics parameters of the rPsRNR.
Table 3. The kinetics parameters of the rPsRNR.
Temperature (°C)Vm (μmol/min/mg)Km (μM)kcat (s−1)
044.9321.00827.126
552.8560.82831.910
1079.2880.79847.867
1584.2500.68450.863
2089.8640.53354.252
25103.690.47262.600
30117.2200.33970.767
Table 4. The thermodynamic parameters of the rPsRNR.
Table 4. The thermodynamic parameters of the rPsRNR.
Temperature (°C)H (KJ/mol)S (J/mol·K)G (KJ/mol)
019.24−146.3059.20
519.20−146.5259.95
1019.16−144.6660.12
1519.12−145.6261.08
2019.08−146.5062.02
2519.03−146.6862.77
3018.99−146.9963.55
Table 5. Primers used for mutations of PsRNR.
Table 5. Primers used for mutations of PsRNR.
MutagenesisPrimers Sequences
T646A5′-CAAAGTTTGTGACGGTAGCTACGACACCATCGAACTC-3′
5′-GAGTTCGATGGTGTCGTAGCTACCGTCACAAACTTTG-3′
F654A5′-CAAATCCGTCAGAGTAATAGCTAAACCAAAGTTTGTGACGG-3′
5′-CCGTCACAAACTTTGGTTTAGCTATTACTCTGACGGATTTG-3′
L665A5′-GTTTGAGATATGCACCGCACCATCGATATACAAATCCG-3′
5′-CGGATTTGTATATCGATGGTGCGGTGCATATCTCAAAC-3′
H667A5′-CACCAACGTTTGAGATAGCCACCAAACCATCGATATAC-3′
5′-GTATATCGATGGTTTGGTGGCTATCTCAAACGTTGGTG-3′

Share and Cite

MDPI and ACS Style

Wang, Y.; Hou, Y.; Nie, P.; Wang, Y.; Ren, X.; Wei, Q.; Wang, Q. A Novel Cold-Adapted and Salt-Tolerant RNase R from Antarctic Sea-Ice Bacterium Psychrobacter sp. ANT206. Molecules 2019, 24, 2229. https://doi.org/10.3390/molecules24122229

AMA Style

Wang Y, Hou Y, Nie P, Wang Y, Ren X, Wei Q, Wang Q. A Novel Cold-Adapted and Salt-Tolerant RNase R from Antarctic Sea-Ice Bacterium Psychrobacter sp. ANT206. Molecules. 2019; 24(12):2229. https://doi.org/10.3390/molecules24122229

Chicago/Turabian Style

Wang, Yatong, Yanhua Hou, Ping Nie, Yifan Wang, Xiulian Ren, Qifeng Wei, and Quanfu Wang. 2019. "A Novel Cold-Adapted and Salt-Tolerant RNase R from Antarctic Sea-Ice Bacterium Psychrobacter sp. ANT206" Molecules 24, no. 12: 2229. https://doi.org/10.3390/molecules24122229

Article Metrics

Back to TopTop