Next Article in Journal
Vibrio Pathogens: A Public Health Concern in Rural Water Resources in Sub-Saharan Africa
Next Article in Special Issue
Selection of a Very Active Microbial Community for the Coupled Treatment of Tetramethylammonium Hydroxide and Photoresist in Aqueous Solutions
Previous Article in Journal
Dietary Inflammatory Index and Cardiometabolic Risk Parameters in Overweight and Sedentary Subjects
Previous Article in Special Issue
A Marine Bacterium, Bacillus sp. Isolated from the Sediment Samples of Algoa Bay in South Africa Produces a Polysaccharide-Bioflocculant
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Novel and Native Microcystin-Degrading Bacterium of Sphingopyxis sp. Isolated from Lake Taihu

Key Laboratory of Environmental Medicine Engineering of Ministry of Education, School of Public Health, Southeast University, Nanjing 210009, China
*
Authors to whom correspondence should be addressed.
Int. J. Environ. Res. Public Health 2017, 14(10), 1187; https://0-doi-org.brum.beds.ac.uk/10.3390/ijerph14101187
Submission received: 22 August 2017 / Revised: 28 September 2017 / Accepted: 30 September 2017 / Published: 6 October 2017
(This article belongs to the Special Issue Microbiological Evaluation of Wastewater Treatment)

Abstract

:
A native, highly efficient microcystin-LR (MC-LR)-degrading bacterium named a7 was isolated from Lake Taihu and identified as Sphingopyxis sp. by 16S rDNA sequence analysis. The strain a7 could totally degrade MC-LR at a rate of 3.33 mg/(L·h), as detected by high-performance liquid chromatography (HPLC). The mlrA, mlrC, and mlrD genes were detected in the strain a7 by sequence analysis. Tetrapeptide and Adda—which are the middle metabolites of MC-LR—were analyzed via liquid chromatography time-of-flight mass spectrometry (LC-TOF-MS) during degradation. These metabolites were degraded completely, which suggested that the native Sphingopyxis sp. a7 was highly efficient in MC-LR degradation under bench conditions. Thus, strain a7 exhibited a significant potential application for bioremediation in water bodies contaminated by MC-LR produced by harmful cyanobacterial blooms.

1. Introduction

Microcystins (MCs) are cyclic heptapeptides with seven amino acids, and are considered one of the most hazardous groups of toxins produced by harmful cyanobacterial blooms (HCBs) [1,2]. These toxins are cyclic heptapeptide hepatotoxins and tumor-promoting substances [3,4]. All MCs consist of the generalized structure cyclo-d-Ala1-X2-d-MeAsp3-Z4-Adda5-d-Glu6-Mdha7 (Figure 1) [5]. MC-LR, which has leucine (L) at position 2 and arginine (R) at position 4, is the most toxic among MCs [6]. Degradation products probably are linearized MC-LR, tetrapeptide, Adda, and other small molecules resulting from cleavage [7,8].
Various aquatic organisms, including molluscs, shrimp, and fish, accumulate MCs in their tissues [9,10,11]. Moreover, algal toxins in drinking water may exceed the standard amount during the bloom, which might increase the risk of cancers [11,12]. With the potential health risks, further studies should be conducted regarding the methods of MC removal.
MC toxins are nonvolatile, hydrophilic, and stable in sunlight over a wide temperature and pH range [13]. Consequently, they are difficult to remove by using the conventional water treatment of chlorine, chlorine dioxide, ozone, and permanganate [14,15]. The specific bacteria appearing along with cyanobacterial blooms can both lower and assimilate the algal toxins and their degradation products [16]. As one of the most important mechanisms for the removal of MCs from the natural environment, microbial degradation is an alternative strategy for physical and chemical water treatment programs [17,18].
Since the first isolation of the bacteria degrading MC in natural waters in 1994, a growing number of MC-degrading bacteria have been isolated in waters of Japan, China, and other parts of the world [19,20,21,22,23,24,25,26].
Lake Taihu, located in the delta of Yangtze River in Eastern China, is the third largest lake in China and offers a resource which supplies water to almost 2 million people for fishery, agriculture, industry, and household drinking. HCBs producing MCs have occurred regularly during the summer in a large part of Lake Taihu in the last three decades [17,22,27].
In previous research from our laboratory, the strain MC-LTH1 of Bordetella sp. and LTH2 of Sarratiamarcesens sp. can totally degrade MC-LR around the rates of 3.0 mg/(L·d) [22,28]. This study aimed to isolate a highly efficient native bacterium that can degrade MC-LR from Lake Taihu, and investigate the possible mechanism of MC degradation to bioremediate the local water environment.

2. Experimental Section

2.1. Collection, Domestication, Isolation, and Culture of MC-Degrading Bacteria

Sludge and water samples were collected from cyanobacteria-salvaged yards in the Guji district in Lake Taihu, China. The sludge (10 g, wet weight) was mixed with 90 mL of phosphate-buffered saline (PBS) by shaking at 30 °C for 30 min at 120 rpm. Subsequently, 0.5 mL of the supernatant was transferred to 4.5 mL of mineral salt medium (MSM) containing crude MC-LR (14.3 mg/L) and incubated with shaking at 30 °C [22]. MC-LR was detected during incubation every 5 days. The mineral salt medium (MSM, pH = 7.0) used for the bacterial isolation and culture contained (g/L) MgSO4·7H2O 1.0, KH2PO4 0.5, K2HPO4 4.0, NaCl 1.0, CaCl2 0.02, FeSO4 0.005, MnCl2·4H2O 0.005, ZnCl2 0.005, and CuCl2 0.0005 [18].
The samples that could degrade MC-LR were selected and treated as follows: MC-LR degradable bacteria taxa was serially diluted (1:10) with MSM containing crude MC-LR from 10−2 to 10−10. These diluted bacteria samples were cultured at 30 °C at 120 rpm for 5 days to degrade MC-LR completely. High-performance liquid chromatography (HPLC) was used to determine the MC-LR concentrations in different bacterial dilutions [28]. Bacterial colonies were selected from the highly efficient degradation sample with the highest dilution.
The selected bacteria (100 μL) were serially-diluted (1:10) with sterile PBS from 10−2 to 10−10. Afterward, 50 μL of diluted solution was spread out on the LB agar plates (2% agar) and cultured for 3 days at 30 °C. Every single colony with differing morphologies was selected and inoculated into MSM containing crude MCs to evaluate their degradation ability [29]. The colonies of MC-degrading bacteria were purified by plate-streaking technology on LB agar plates [30]. The bacterial culture and algal toxin liquid plate were crossed to determine the inbred strains. The inbred strains were transferred for the second verification.

2.2. Identification of 16S rDNA from MC-Degrading Bacteria

2.2.1. Amplification of 16S rDNA

DNA extraction kit (Tiangen, Beijing, China) was used to extract the DNA of MC-degrading bacteria. The PCR reaction system was 50 μL, including 100 ng of template DNA, 1 × PCR Buffer (Mg2+ Free), 1.5 mM MgCl2, 0.2 mM dNTP mixture, 1.25 units of Taq DNA polymerase, and 400 nM of each of the following primers: 27F 5′-AGAGTTTGATCMTGGCTCAG-3′, 1492R 5′-TACGGYTACCTTGTTACGACTT-3′ [31]. The primers were synthesized by Shanghai Invitrogen.
The PCR condition included the initial denaturation at 98 °C for 5 min; 94 °C for 30 s, annealing for 30 s at 60 °C, 30 cycles of extension for 90 s at 72 °C; and a final elongation for 10 min at 72 °C. The PCR products were electrophoresed by agar gel with 2% agarose at 100 V for 40 min [31].

2.2.2. Determination of 16S rDNA Sequence and Construction of Phylogenetic Tree

The PCR products of 16S rDNA were sent to the BGI Company (Shanghai, China) in 4 °C then sequenced. The sequencing results were submitted to the United States National Center for Biotechnology Information Website (NCBI) to obtain the corresponding registration number. Furthermore, nucleotide sequences were searched in GenBank for BLAST comparison, and the homology was analyzed. Sequence similarity search was conducted using the NCBI BLAST (https://blast.ncbi.nlm.nih.gov/Blast.cgi) network service [29]. Afterward, multiple sequence alignment and a phylogenetic tree were constructed using the software MEGA5 [32].

2.3. Cloning of Degrading Enzyme Genes from MC-Degrading Bacteria a7

mlrA, mlrB, mlrC, and mlrD specific primers were used to amplify the DNA of strain a7 according to the literature. The primer sequences are shown in Table 1. The PCR reaction system was 50 μL, which included 100 ng of template DNA, 1 × PCR buffer (Mg2+ Free), 1.5 mM MgCl2, 0.2 mM dNTP mixture, 0.4 mM of each primer, and 1.25 units of Taq DNA polymerase. The PCR condition included the initial denaturation at 95 °C for 5 min; 98 °C for 10 s, annealing for 30 s at 56 °C, 30 cycles of extension for 30 s at 72 °C; and a final elongation for 5 min at 72 °C and held at 4 °C. The PCR products were detected by the agarose gel electrophoresis and subsequently sequenced same with 16S rDNA.
The amino acid sequences of degrading enzyme genes were translated by the nucleotide sequences using the Sixpack program of EMBOSS (http://www.ebi.ac.uk/Tools/st/) [25]. The mlr genes were blasted by BLASTx program on the NCBI, and corresponding amino acid sequences were retrieved. Amino acid phylogenetic trees were constructed by MEGA5 software using the neighbor-joining algorithm based on distances calculated by Clustalw method [25].

2.4. Collection, Domestication, Isolation, and Culture of MC-Degrading Bacteria

The standard MC-LR was diluted with deionized water and MSM. The initial concentration of MC-LR was 20 mg/L. To investigate the degradation rate of standard MC-LR, the bacterial strain a7 was cultured on an orbital shaker at 120 rpm in MC-LR at 30 °C for 48 h in MSM containing crude MCs. The bacterial cells were centrifuged (5000× g, 15 min, 4 °C) and washed with PBS (Beyotime, Shanghai, China) twice. The bacterial strain a7 which was resuspended into MSM containing standard MC-LR was cultured at a constant condition (120 rpm, 30 °C) and analyzed by HPLC every 1 h [22].
MC-LR was measured by HPLC system (Agilent 1100, Santa Clara, CA, USA) with a Zorbax Extend C18 column (4.6 × 50 mm, 5 μm, Agilent, Santa Clara, CA, USA), which was kept at 40 °C. A mixture of methanol and 0.05% trifluoroacetic acid aqueous solution (53:47, v/v) was used as mobile phase. The flow rate was 1 mL/min, and the injection volume was 20 μL with 238 nm of variable wavelength detector [28].
The main degradation products of MC-LR were identified by a liquid chromatography time-of-flight mass spectrometry (LC-TOF-MS, Agilent 1200-6224, Santa Clara, CA,) in positive electrospray ionization (ESI) mode. A Zorbax Extend C18 column (2.1 × 50 mm, 1.8 μm, Agilent, Santa Clara, CA, USA) was used at 40 °C, and the injection volume was 2 μL. The mobile phase was a mixture of methanol and 0.1% formic acid aqueous solution (55:45, v/v) at a flow rate of 0.08 mL/min. The ESI-TOF-MS parameters were as follows: capillary voltage, 3500 V; gas temperature, 325 °C; flow of drying gas, 9 L/min; nebulizer pressure, 40 psi; and fragmentor voltage, 175 V. The flow rate was 0.1 mL/min [28].

3. Results

3.1. Isolation of MC-LR-Degrading Bacterium

According to the process of collection, domestication, isolation, and culture of MC-degrading bacteria, only one bacterium strain with high MC-degrading activities, named as a7, was obtained. In the selection of degrading bacteria, eight different colony types could survive in MC-LR media but could not degrade MC-LR.

3.2. Identification of MC-Degrading Bacterium by 16S rDNA Sequence Analysis

The PCR products of the 16S rDNA were sequenced by the BGI Company, and the accession number in the NCBI is KU954525. The phylogenetic tree was constructed with the homology of 97%. The nucleotide sequence of 16S rDNA from a7 was mostly similar to that of Sphingopyxis macrogoltabida strain NBRC 15033 (99% similarity, accession number: NR_113720.1). It suggested that strain a7 belongs to genus Sphingopyxis. The phylogenetic tree was constructed using a neighbor-joining analysis with 1000 bootstrap replication, as shown in Figure 2.

3.3. Cloning and Sequencing of the Key Genes of Degrading Enzyme from MC-Degrading Bacterium a7

The mlrA, mlrB, mlrC, and mlrD specific primers were used for PCR amplification of a7, and the agarose gel electrophoresis of the PCR products is shown in Figure 3. Three bright DNA bands were observed after amplification, except for mlrB gene, and the fragment size was about 750, 1300, and 1000 bp. Thus, strain a7 contained the key enzyme gene mlrA, mlrC, and mlrD, which might be involved in the process of MC degradation. The PCR fragments were subsequently sequenced. The sequences of mlrA, mlrC, and mlrD from a7 were 99%, 100%, and 98% similar to the genes obtained from Sphingomonas sp. USTB-05, Sphingomonas sp. USTB-05, and Sphingopyxis sp. C-1, respectively. The amino acid phylogenetic trees of mlrA, mlrC, and mlrD genes were constructed based on the comparison with amino acid sequences of high-homology bacteria (MEGA5, Figure 4).

3.4. Biodegradation of MC-LR by Strain a7 and Analysis of Main Degradation Products

About 20 mg/L MC-LR was completely degraded by strain a7 within 6 h, and the average degradation rate was 3.33 mg/(L·h) (Figure 5). MC-LR was degraded immediately without any lag phase, and more than 80% MC-LR was decomposed in the first two hours. Moreover, with the increase of time, degradation rate was progressively slower and eventually halted at 6h. Two major intermediate degradation products—namely, peaks A and B—were detected in 6 h using HPLC (Figure 6b). Strain a7 decomposed peaks A and B completely in 12 h (Figure 6c).
Peaks A and B in 6 h samples were further detected by LC-TOF-MS (Figure 7). The intermediate degradation product B exhibited a protonated molecular ion at m/z 615.3398 (C32H46N4O8), which was considered as tetrapeptide (Adda-Glu-Mdha-Ala) (Figure 7a). The intermediate product A exhibited several accompanying ions at m/z 315.1955 ([M+H−NH3]+) and 283.1700 ([M+H−NH3−MeOH]+), which were identified as Adda and its fragment, respectively (Figure 7b).

4. Discussion

In this study, an efficient and novel MC-degrading bacterium a7 was isolated from the bacterial communities of a lake. Moreover, 16srDNA of this a7 was characterized and identified as Sphingopyxis. Sphingopyxis sp. is a group of homogeneous microorganisms in the α-Proteobacteria class and has been isolated from nature samples before [16,19,26,35,36]. Sphingopyxis can tolerate the conditions of extreme poor nutrition, utilize various simple molecules, and decompose complex organic matter [26].
The degrading rate of MC by a7 was around 3.33 mg/(L·h). Compared with that of the previous studies, the degrading rate was six times greater than Sphingomonas sp. ACM-3962, 14 times of Sphingomonas sp. Y2, and hundredfolds of Sphingopyxis sp. LH21 which were designated as notable microorganisms with highly efficient of MC-degradation from natural waterbody [16,24,37,38]. The degradation rates varied from 0.06 μg/(L·d) to 101.52 mg/(L·d) in different research studies, probably caused by the different bacterial strains or other physical and chemical conditions utilized in the laboratory. Some studies have reported that MC-degrading bacteria had the most efficient degradation activity within 25–30 °C. The degrading enzymes were most probably inhibited when temperature was up 30 °C or below 25 °C [37,39,40]. So, in this study, the degrading rate was tested on 30 °C.
The discovery of bacteria that degrade MC-LR may indicate that these functional bacteria are prevalent in the natural environment [41]. Application of bacteria which are capable of decreasing microcystins dramatically during the outbreak of cyanobacterial-dominated harmful algal blooms is considered to be an important biological method toward diminishing the potential adverse effects of microcystins [28]. Table 2 shows the MC-degrading rates of bacteria isolated in individual research and different areas in the world.
The MC-degrading capability was attributed to mlrA, mlrB, mlrC, and mlrD genes in the bacteria reported by Kormas and et al. [6,7,25,37]. Based on the results of the presented study, microcystinase (MlrA) hydrolyzes cyclic MC-LR into a linear intermediate (Adda-Glu-Mdha-Ala-Leu-MeAsp-Arg) by undermining the peptide bond connecting Adda and arginine [25]. Further decomposition of linearized MC-LR by MlrB to generated a tetrapeptide (Adda-Glu-Mdha-Ala, 614, C32H46N4O8) [16]. MlrC exhibited a double-degrading activity, which was decomposing both linear MC-LR and the tetrapeptide into Adda (331, C20H29NO3) and other small molecular substances [7]. MlrD is predicted to transport MC-LR and its degradation products across the bacterial cell wall [37]. In our study, the ion of degradation product B presented a similar molecular weight (615.3398) with tetrapeptide. Therefore, degradation product B was determined as tetrapeptide. The protonated ion A (m/z 332.2208) was consistent with Adda. Ions at m/z 315.1955 (consistent with [adda+H-NH3]+) and 283.1700 (consistent with [adda+H−NH3−MeOH]+) were fragments of Adda. So, the degradation product A can be determined as Adda generally. Furthermore, mlrA, mlrC, and mlrD were detected in Sphingopyxis a7. Therefore, the possible degradation pathway of MC-LR might be shown in Figure 8 and MC-LR can be degraded by Sphingopyxis sp. a7 completely. To our knowledge, this study is the first report that MC-LR and Adda which is the main active structure of MC-LR can be decomposed simultaneously using single bacterial strain Sphingopyxis sp. [7,16,34]. MC-degrading and detoxication properties of strain a7 can potentially be applied for bioremediation in the harmful cyanobacterial blooms and contribute to improve water quality.

5. Conclusions

A native and highly efficient MC-LR-degrading bacterium was isolated from Lake Taihu and identified as Sphingopyxis sp., according to the 16S rDNA sequence. The MC-LR degrading rate was 3.33 mg/(L·h) at bench conditions. Strain a7 contained the key enzyme genes mlrA, mlrC, and mlrD, which might be involved in MC degradation. The degradation products included tetrapeptide and Adda, which were the metabolites of MC-LR catalyzed by microcystinases. This is the first report finding that MC-LR and Adda can be decomposed totally by Sphingopyxis sp.
Therefore, isolated Sphingopyxis sp. a7 may be used to solve the problem of MC-LR pollution. Further studies should apply a7 to an actual environmental bioremediation project.

Acknowledgments

This work was supported by the National Science and Technology Major Project (Grant No. 2012ZX07101-005).

Author Contributions

Yuepu Pu and Juan Zhang designed the study, Qingqing Lu, and Qin Ding conducted the experiments and collected the data, Juan Zhang wrote the paper. Lihong Yin and Yuepu Pu critically revised the manuscript. All authors participated in the preparation of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dawson, R.M. The toxicology of microcystins. Toxicon 1998, 36, 953–962. [Google Scholar] [CrossRef]
  2. Agha, R.; Cires, S.; Wormer, L.; Quesada, A. Limited stability of microcystins in oligopeptide compositions of Microcystis aeruginosa (Cyanobacteria): Implications in the definition of chemotypes. Toxins 2013, 5, 1089–1104. [Google Scholar] [CrossRef] [PubMed]
  3. Falconer, I.R. Quality and Treatment of Drinking Water II: The Handbook of Environmental Chemistry; Springer: Berlin, Germany, 1998; pp. 53–82. [Google Scholar]
  4. Humpage, A.R.; Hardy, S.J.; Moore, E.J.; Froscio, S.M.; Falconer, I.R. Microcystins (Cyanobacterial toxins) in drinking water enhance the growth of aberrant crypt foci in the mouse colon. J. Toxicol. Environ. Health A 2000, 61, 155–165. [Google Scholar] [PubMed]
  5. Codd, G.; Bell, S.; Kaya, K.; Ward, C.; Beattie, K.; Metcalf, J. Cyanobacterial toxins, exposure routes and human health. Eur. J. Phycol. 1999, 34, 405–415. [Google Scholar] [CrossRef]
  6. Imanishi, S.; Kato, H.; Mizuno, M.; Tsuji, K.; Harada, K. Bacterial degradation of Microcystins and Nodularin. Chem. Res. Toxicol. 2005, 18, 591–598. [Google Scholar] [CrossRef] [PubMed]
  7. Shimizu, K.; Maseda, H.; Okano, K.; Kurashima, T.; Kawauchi, Y.; Xue, Q.; Utsumi, M.; Zhang, Z.; Sugiura, N. Enzymatic pathway for biodegrading Microcystin LR in Sphingopyxis sp. C-1. J. Biosci. Bioeng. 2012, 114, 630–634. [Google Scholar] [CrossRef] [PubMed]
  8. Hashimoto, E.H.; Kato, H.; Kawasaki, Y.; Nozawa, Y.; Tsuji, K.; Hirooka, E.Y.; Harada, K. Further investigation of microbial degradation of microcystin using the advanced Marfey method. Chem. Res. Toxicol. 2009, 22, 391–398. [Google Scholar] [CrossRef] [PubMed]
  9. Gurbuz, F.; Uzunmehmetoglu, O.Y.; Diler, O.; Metcalf, J.S.; Codd, G.A. Occurrence of Microcystins in water, bloom, sediment and fish from a public water supply. Sci. Total Environ. 2016, 562, 860–868. [Google Scholar] [CrossRef] [PubMed]
  10. Paerl, H.W.; Huisman, J. Climate change: A catalyst for global expansion of harmful Cyanobacterial blooms. Environ. Microbiol. Rep. 2009, 1, 27–37. [Google Scholar] [CrossRef] [PubMed]
  11. Yu, S.Z. Drinking water and primary liver cancer. In Primary Liver Cancer; Tang, Z.Y., Wu, M.C., Xia, S.S., Eds.; Springer-Verlag: Beijing, China, 1989; pp. 30–37. [Google Scholar]
  12. Ueno, Y.; Nagata, S.; Tsutsumi, T.; Hasegawa, A.; Watanabe, M.F.; Park, H.D.; Chen, G.C.; Chen, G.; Yu, S.Z. Detection of Microcystins, a blue-green algal Hepatotoxin, in drinking water sampled in Haimen and Fusui, endemic areas of primary liver cancer in China, by highly sensitive immunoassay. Carcinogenesis 1996, 17, 1317–1321. [Google Scholar] [CrossRef] [PubMed]
  13. World Health Organization. Cyanobacterial Toxins: Microcystin-LR in Drinking-Water. Background Document for Development of WHO Guidelines for Drinking-Water Quality, 2nd ed.; World Health Organization: Geneva, Switzerland, 2003. [Google Scholar]
  14. Lawton, L.A.; Robertson, P. Physico-chemical treatment methods for the removal of Microcystins (Cyanobacterial Hepatotoxins) from potable waters. Chem. Soc. Rev. 1999, 28, 217–224. [Google Scholar] [CrossRef]
  15. Saitou, T.; Sugiura, N.; Itayama, T.; Inamori, Y.; Matsumura, M. Degradation characteristics of Microcystins by isolated bacteria from Lake Kasumigaura. J. Water Supply Res. Technol.–AQUA 2003, 52, 13–18. [Google Scholar]
  16. Bourne, D.G.; Jones, G.J.; Blakeley, R.L.; Jones, A.; Negri, A.P.; Riddles, P. Enzymatic pathway for the bacterial degradation of the Cyanobacterial cyclic peptide toxin Microcystin LR. Appl. Environ. Microbiol. 1996, 62, 4086–4094. [Google Scholar] [PubMed]
  17. Chen, J.; Hu, L.B.; Zhou, W.; Yan, S.H.; Yang, J.D.; Xue, Y.F.; Shi, Z.Q. Degradation of Microcystin-LR and RR by a Stenotrophomonas sp. strain EMS isolated from Lake Taihu, China. Int. J. Mol. Sci. 2010, 11, 896–911. [Google Scholar] [CrossRef] [PubMed]
  18. Gagala, I.; Mankiewicz-Boczek, J. The natural degradation of Microcystins (Cyanobacterial Hepatotoxins) in fresh water—The future of modern treatment systems and water quality improvement. Pol. J. Environ. Stud. 2012, 21, 1125–1139. [Google Scholar]
  19. Jones, G.J.; Bourne, D.G.; Blakeley, R.L.; Doelle, H. Degradation of the Cyanobacterial Hepatotoxin Microcystin by aquatic bacteria. Nat. Toxins 1994, 2, 228–235. [Google Scholar] [CrossRef] [PubMed]
  20. Okano, K.; Shimizu, K.; Kawauchi, Y.; Maseda, H.; Utsumi, M.; Zhang, Z.; Neilan, B.A.; Sugiura, N. Characteristics of a Microcystin-degrading bacterium under Alkaline environmental conditions. J. Toxicol. 2009, 2009, 1–8. [Google Scholar] [CrossRef] [PubMed]
  21. Kazuya, S.; Kunihiro, O.; Hideaki, M.; Tomoaki, I.; Yukio, K. Finding of Microcystin-Degrading Bacterium and Elucidation of Its Degradation Mechanism. Available online: https://www.pref.ibaraki.jp/soshiki/seikatsukankyo/kasumigauraesc/04_kenkyu/kaigi/docments/kosyou/13/2009wlc_SU.pdf (accessed on 28 September 2017).
  22. Yang, F.; Zhou, Y.; Sun, R.; Wei, H.; Li, Y.; Yin, L.; Pu, Y. Biodegradation of microcystin-LR and-RR by a novel microcystin-degrading bacterium isolated from Lake Taihu. Biodegradation 2014, 25, 447–457. [Google Scholar] [CrossRef] [PubMed]
  23. Ho, L.; Hoefel, D.; Saint, C.P.; Newcombe, G. Isolation and identification of a novel Microcystin-degrading bacterium from a biological sand filter. Water Res. 2007, 41, 4685–4695. [Google Scholar] [CrossRef] [PubMed]
  24. Phujomjai, Y.; Somdee, A.; Somdee, T. Biodegradation of microcystin [Dha(7)]MC-LR by a novel microcystin-degrading bacterium in an internal airlift loop bioreactor. Water Sci. Technol. 2016, 73, 267–274. [Google Scholar] [CrossRef] [PubMed]
  25. Kormas, K.A.; Lymperopoulou, D.S. Cyanobacterial Toxin Degrading Bacteria: Who Are They? Biomed. Res. Int. 2013, 6, 463894. [Google Scholar] [CrossRef] [PubMed]
  26. Berg, K.A.; Lyra, C.; Sivonen, K.; Paulin, L.; Suomalainen, S.; Tuomi, P.; Rapala, J. High diversity of cultivable heterotrophic bacteria in association with Cyanobacterial water blooms. ISME J. 2009, 3, 314–325. [Google Scholar] [CrossRef] [PubMed]
  27. Zhang, D.; Xie, P.; Liu, Y.; Chen, J.; Liang, G. Bioaccumulation of the hepatotoxic Microcystins in various organs of a freshwater snail from a subtropical Chinese lake, Taihu Lake, with dense toxic Microcystis blooms. Environ. Toxicol. Chem. 2007, 26, 171–176. [Google Scholar] [CrossRef] [PubMed]
  28. Yang, F.; Zhou, Y.; Yin, L.; Zhu, G.; Liang, G.; Pu, Y. Microcystin-degrading activity of an indigenous bacterial strain Stenotrophomonas acidaminiphila MC-LTH2 isolated from Lake Taihu. PLoS ONE 2014, 9, e86216. [Google Scholar] [CrossRef] [PubMed]
  29. Li, H.; Ai, H.N.; Kang, L.; Sun, X.F.; He, Q. Simultaneous Microcystis algicidal and Microcystin degrading capability by a single Acinetobacter bacterial strain. Environ. Sci. Technol. 2016, 50, 11903–11911. [Google Scholar] [CrossRef] [PubMed]
  30. Somdee, T.; Thunders, M.; Ruck, J.; Lys, I.; Allison, M.; Page, R. Degradation of [Dha(7)]MC-LR by a microcystin degrading bacterium isolated from Lake Rotoiti, New Zealand. ISRN Microbiol. 2013, 2013, 1–8. [Google Scholar] [CrossRef] [PubMed]
  31. Yang, F.; Wei, H.Y.; Li, X.Q.; Li, Y.H.; Li, X.B.; Yin, L.H.; Pu, Y.P. Isolation and characterization of an algicidal bacterium indigenous to lake Taihu with a red pigment able to lyse Microcystis aeruginosa. Biomed. Environ. Sci. 2013, 26, 148–154. [Google Scholar] [PubMed]
  32. Tamura, K.; Peterson, D.; Peterson, N.; Stecher, G.; Nei, M.; Kumar, S. MEGA5: Molecular evolutionary genetics analysis using maximum likelihood, evolutionary distance, and maximum parsimony methods. Mol. Biol. Evol. 2011, 28, 2731–2739. [Google Scholar] [CrossRef] [PubMed]
  33. Saito, T.; Okano, K.; Park, H.D.; Itayama, T.; Inamori, Y.; Neilan, B.A.; Burns, B.P.; Sugiura, N. Detection and sequencing of the microcystin LR-degrading gene, mlrA, from new bacteria isolated from Japanese lakes. FEMS Microbiol. Lett. 2003, 229, 271–276. [Google Scholar] [CrossRef]
  34. Shimizu, K.; Maseda, H.; Okano, K.; Itayama, T.; Kawauchi, Y.; Chen, R.; Utsumi, M.; Zhang, Z.; Sugiura, N. How Microcystin-degrading bacteria express Microcystin degradation activity. Lakes Reserv. Res. Manag. 2011, 16, 169–178. [Google Scholar] [CrossRef]
  35. Xu, H.; Wang, H.; Xu, Q.; Lv, L.; Yin, C.; Liu, X.; Du, H.; Yan, H. Pathway for biodegrading microcystin-YR by Sphingopyxis sp. USTB-05. PLoS ONE 2015, 10, e124425. [Google Scholar] [CrossRef] [PubMed]
  36. Zhang, M.; Pan, G.; Yan, H. Microbial biodegradation of Microcystin-RR by bacterium Sphingopyxis sp. USTB-05. J. Environ. Sci. 2010, 22, 168–175. [Google Scholar] [CrossRef]
  37. Bourne, D.G.; Riddles, P.; Jones, G.J.; Smith, W.; Blakeley, R.L. Characterisation of a gene cluster involved in bacterial degradation of the Cyanobacterial toxin Microcystin LR. Environ. Toxicol. 2001, 16, 523–534. [Google Scholar] [CrossRef] [PubMed]
  38. Jones, G.J.; Orr, P.T. Release and degradation of Microcystin following algicide treatment of a Microcystis-aeruginosa bloom in a recreational lake, as determined by HPLC and protein phosphatase inhibition assay. Water Res. 1994, 28, 871–876. [Google Scholar] [CrossRef]
  39. Chen, X.G.; Yang, X.; Yang, L.L.; Xiao, B.D.; Wu, X.Q.; Wang, J.T.; Wan, H.G. An effective pathway for the removal of Microcystin LR via anoxic biodegradation in lake sediment. Water Res. 2010, 44, 1884–1892. [Google Scholar] [CrossRef] [PubMed]
  40. Ho, L.; Hoefel, D.; Palazot, S.; Sawade, E.; Newcombe, G.; Saint, C.P.; Brooks, J.D. Investigations into the biodegradation of Microcystin-LR in wastewaters. J. Hazard. Mater. 2010, 180, 628–633. [Google Scholar] [CrossRef] [PubMed]
  41. Eleuterio, L.; Batista, J.R. Biodegradation studies and sequencing of Microcystin-LR degrading bacteria isolated from a drinking water biofilter and a fresh water lake. Toxicon 2010, 55, 1434–1442. [Google Scholar] [CrossRef] [PubMed]
  42. Yan, H.; Wang, J.; Chen, J.; Wei, W.; Wang, H.; Wang, H. Characterization of the first step involved in enzymatic pathway for Microcystin-RR biodegraded by Sphingopyxis sp. USTB-05. Chemosphere 2012, 87, 12–18. [Google Scholar] [CrossRef] [PubMed]
  43. Valeria, A.M.; Ricardo, E.J.; Stephan, P.; Alberto, W.D. Degradation of Microcystin-RR by Sphingomonas sp. CBA4 isolated from San Roque reservoir (Córdoba–Argentina). Biodegradation 2006, 17, 447–455. [Google Scholar] [CrossRef] [PubMed]
  44. Ishii, H.; Nishijima, M.; Abe, T. Characterization of degradation process of Cyanobacterial Hepatotoxins by a gram-negative aerobic bacterium. Water Res. 2004, 38, 2667–2676. [Google Scholar] [CrossRef] [PubMed]
  45. Tsuji, K.; Asakawa, M.; Anzai, Y.; Sumino, T.; Harada, K. Degradation of Microcystins using immobilized microorganism isolated in an eutrophic lake. Chemosphere 2006, 65, 117–124. [Google Scholar] [CrossRef] [PubMed]
  46. Jiang, Y.; Shao, J.; Wu, X.; Xu, Y.; Li, R. Active and silent members in the mlr gene cluster of a Microcystin-degrading bacterium isolated from Lake Taihu, China. FEMS Microbiol. Lett. 2011, 322, 108–114. [Google Scholar] [CrossRef] [PubMed]
  47. Hu, L.B.; Yang, J.D.; Zhou, W.; Yin, Y.F.; Chen, J.; Shi, Z.Q. Isolation of a Methylobacillus sp. that degrades microcystin toxins associated with cyanobacteria. New Biotechnol. 2009, 26, 205–211. [Google Scholar] [CrossRef] [PubMed]
  48. Lemes, G.F.; Kersanach, R.; Da, S.; Pinto, L.; Dellagostin, O.A.; Yunes, J.S.; Matthiensen, A. Biodegradation of microcystins by aquatic Burkholderia sp. from a South Brazilian coastal lagoon. Ecotoxicol. Environ. Saf. 2008, 69, 358–365. [Google Scholar] [CrossRef] [PubMed]
  49. Lemes, G.; Kist, L.W.; Bogo, M.R.; Yunes, J.S. Biodegradation of [D-Leu(1)]microcystin-LR by a bacterium isolated from sediment of Patos Lagoon estuary, Brazil. J. Venomanim. Toxins 2015, 21, 1–8. [Google Scholar]
  50. Dziga, D.; Lisznianska, M.; Wladyka, B. Bioreactor study employing bacteria with enhanced activity toward Cyanobacterial toxins Microcystins. Toxins 2014, 6, 2379–2392. [Google Scholar] [CrossRef] [PubMed]
  51. Takenaka, S.; Watanabe, M.F. Microcystin LR degradation by Pseudomonas aeruginosa alkaline protease. Chemosphere 1997, 34, 749–757. [Google Scholar] [CrossRef]
  52. Zhang, J.; Shi, H.; Liu, A.; Cao, Z.; Hao, J.; Gong, R. Identification of a new Microcystin-degrading bacterium isolated from lake Chaohu, China. Bull. Environ. Contam. Toxicol. 2015, 94, 661–666. [Google Scholar] [CrossRef] [PubMed]
  53. Hu, L.; Zhang, F.; Liu, C.; Wang, M. Biodegradation of Microcystins by Bacillus sp. strain EMB. Energy Procedia 2012, 16, 2054–2059. [Google Scholar] [CrossRef]
  54. Nybom, S.K.; Dziga, D.; Heikkilä, J.E.; Kull, T.J.; Salminen, S.J.; Meriluoto, J.O. Characterization of Microcystin-LR removal process in the presence of probiotic bacteria. Toxicon 2012, 59, 171–181. [Google Scholar] [CrossRef] [PubMed]
  55. Alamri, S.A. Biodegradation of Microcystin-RR by Bacillus flexus isolated from a Saudi freshwater lake. Saudi J. Biol. Sci. 2012, 19, 435–440. [Google Scholar] [CrossRef] [PubMed]
  56. Ramani, A.; Rein, K.; Shetty, K.G.; Jayachandran, K. Microbial degradation of Microcystin in Florida’s freshwaters. Biodegradation 2012, 23, 35–45. [Google Scholar] [CrossRef] [PubMed]
  57. Manage, P.M.; Edwards, C.; Singh, B.K.; Lawton, L.A. Isolation and identification of novel microcystin-degrading bacteria. Appl. Environ. Microbiol. 2009, 75, 6924–6928. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Chemical structure of microcystins (MCs). X and Z are variable l-amino acids [5].
Figure 1. Chemical structure of microcystins (MCs). X and Z are variable l-amino acids [5].
Ijerph 14 01187 g001
Figure 2. A phylogenetic tree based on the bacterial 16S rDNA sequence from the isolated strain a7 and the closely related species. The numbers at the nodes are the levels of bootstrap support (%) based on the neighbor-joining analyses of 1000 resampled datasets. The scale bar represents 0.02 nucleotide substitutions per position.
Figure 2. A phylogenetic tree based on the bacterial 16S rDNA sequence from the isolated strain a7 and the closely related species. The numbers at the nodes are the levels of bootstrap support (%) based on the neighbor-joining analyses of 1000 resampled datasets. The scale bar represents 0.02 nucleotide substitutions per position.
Ijerph 14 01187 g002
Figure 3. mlrA, mlrB, mlrC and mlrD PCR amplified products gel electrophoresis of strain a7.
Figure 3. mlrA, mlrB, mlrC and mlrD PCR amplified products gel electrophoresis of strain a7.
Ijerph 14 01187 g003
Figure 4. Amino acid sequences phylogenetic trees of mlrA, mlrC, and mlrD. The numbers at the nodes are the levels of bootstrap support (%) based on the neighbor-joining analyses of 1000 resampled datasets. The scale bar represents 0.01 estimated distance per position. (a) mlrA gene; (b) mlrC gene; (c) mlrD gene.
Figure 4. Amino acid sequences phylogenetic trees of mlrA, mlrC, and mlrD. The numbers at the nodes are the levels of bootstrap support (%) based on the neighbor-joining analyses of 1000 resampled datasets. The scale bar represents 0.01 estimated distance per position. (a) mlrA gene; (b) mlrC gene; (c) mlrD gene.
Ijerph 14 01187 g004
Figure 5. Degradation curve of standard MC-LR using strain a7.
Figure 5. Degradation curve of standard MC-LR using strain a7.
Ijerph 14 01187 g005
Figure 6. Degradation of standard microcystin LR using strain a7. High-performance liquid chromatography chromatogram at (a) 0, (b) 6 h, and (c) 12 h. Peaks A and B show two intermediate products of MC-LR.
Figure 6. Degradation of standard microcystin LR using strain a7. High-performance liquid chromatography chromatogram at (a) 0, (b) 6 h, and (c) 12 h. Peaks A and B show two intermediate products of MC-LR.
Ijerph 14 01187 g006
Figure 7. Electrospray ionization time-of-flight mass spectrometry (ESI-TOF-MS) spectrum of MC-LR degradation product. (a) MC-LR degradation product peak B with a molecular weight of 615.3398 ([M+H]+). (b) MC-LR degradation product peak A with a molecular weight of 283.1700 ([M+H−NH3−MeOH]+).
Figure 7. Electrospray ionization time-of-flight mass spectrometry (ESI-TOF-MS) spectrum of MC-LR degradation product. (a) MC-LR degradation product peak B with a molecular weight of 615.3398 ([M+H]+). (b) MC-LR degradation product peak A with a molecular weight of 283.1700 ([M+H−NH3−MeOH]+).
Ijerph 14 01187 g007
Figure 8. Possible degradation pathway of MC-LR by strain a7. The position of cleavage is marked with an arrow (↓).
Figure 8. Possible degradation pathway of MC-LR by strain a7. The position of cleavage is marked with an arrow (↓).
Ijerph 14 01187 g008
Table 1. Specific primer sequences.
Table 1. Specific primer sequences.
PrimersSequenceReferences
mlrA-Forward5′-GACCCGATGTTCAAGATACT-3′[33]
mlrA-Reverse5′-CTCCTCCCACAAATCAGGAC-3′
mlrB-Forward5′-CTCGATGCGGTATTTGCTG-3′[34]
mlrB-Reverse5′-TCCAACGACCATCCCTTCTG-3′
mlrC-Forward5′-CGAAGGCGAAAGGTGCAAC-3′[35]
mlrC-Reverse5′-GAGCGCTTGTGATAGTGACG-3′
mlrD-Forward5′-GTTCCTCGGCGTAGCCT-3′[24]
mlrD-Reverse5′-GCGACGAAGATCGTTGCT-3′
Table 2. Microcystin-degrading bacteria isolated in different areas and degradation rates.
Table 2. Microcystin-degrading bacteria isolated in different areas and degradation rates.
BacteriaSource16S rRNA GenBank Accession NumberDegradable Microcystin AnaloguesDegradation Rate μg/(L·d)References
α-Proteobacteria
Sphingopyxis sp. USTB-05-EF607053MC-RR; MC-YR101,520; 35,520[35,36,42]
Sphingopyxis sp. LH21Torrens Lake in South AustraliaDQ112242MC-LR; MC-LA2.08; 2.50[24]
Sphingomonas sp. CBA4San Roque reservoirAY920497MC-RR130[43]
Sphingomonas sp. 7CYSurface water sample of Lake SuwaAB076083MC-LR; MC-RR1500; 1000[44]
Sphingomonas sp. Y2A hypertrophic lakeNR_040927/AB084247MC-LR; MC-RR; MC-YR5400; 13,000; -[37]
Sphingomonas sp. ACM-3962Surface water, AustraliaAF411072MC-LR; MC-RR12,100 #; 12,100 #[16,38]
Sphingomonas sp. B9Tsukui lakeAB159609MC-LR; MC-RR25; 75[45]
Sphingomonas stygiaNatural lakes-MC-LR; MC-RR; MC-YR660; 1900; 1250[15]
Novosphingobium sp. THN1Lake TaihuHQ664117MC-LR2700[46]
β-Proteobacteria
Bordetella sp. MCYF11Lake TaihuKC734882MC-LR; MC-RR7440;4080[22]
Methylobacillus sp. J10Lake TaihuFJ4185MC-LR; MC-RR4940; 6350[47]
Burkholderia sp. UPC-BI05A South Brazilian coastal lagoon surface waterDQ459360MC-LR50[48]
γ-Proteobacteria
Pseudomonas aeruginosa. DMxSPatos Lagoon estuaryHQ890467MC-LR0.06[49]
Stenotrophomonas acidaminiphila YF-S2Lake TaihuKF305533MC-LR; MC-RR3000; 5600[28]
Escherichia coli--MC-LR5270[50]
Pseudomonas aeruginosaReservoir surface water-MC-LR2500[51]
Morganella morganii C25216Las Vegas Bay of Lake Mead-MC-LR4.75[42]
Stenotrophomonas sp. EMSLake Taihu-MC-LR; MC-RR700; 1700[30]
Firmicutes
Bacillus sp. JZ-2013Lake ChaohuKF841622MC-LR1330[52]
Bacillus sp. EMBLake TaihuFJ526332MC-LR; MC-RR2150; 2990[53]
Lactobacillus rhamnosus GG and LC-705Probiotic strains-MC-LR65 and 55[54]
Bacillus flexus SSZ01a Saudi eutrophic lake-MC-RR2500[55]
Actinobacteria
Microbacterium sp. DC8Lake Okeechobee-MC-LR8.60[56]
Bifidobacterium longum 46Probiotic strains-MC-LR70[54]
Rhodococcus sp. C1,C3Scottish water bodiesFN392688, FN392689MC-LR4,000[57]
Arthrobacter sp. C6, F7, F10, R1, R1, R6 and R9Scottish water bodiesFN392690, FN392692-FN392697MC-LR5000, 5000, 4000, 5000, 5000, 5000, and 5000[57]
Brevibacterium sp. F10Scottish water bodiesFN392691MC-LR5000[57]
-: no detection /no mention. #: Degradation rate of this strain calculated by Eleuterio and Batista.

Share and Cite

MDPI and ACS Style

Zhang, J.; Lu, Q.; Ding, Q.; Yin, L.; Pu, Y. A Novel and Native Microcystin-Degrading Bacterium of Sphingopyxis sp. Isolated from Lake Taihu. Int. J. Environ. Res. Public Health 2017, 14, 1187. https://0-doi-org.brum.beds.ac.uk/10.3390/ijerph14101187

AMA Style

Zhang J, Lu Q, Ding Q, Yin L, Pu Y. A Novel and Native Microcystin-Degrading Bacterium of Sphingopyxis sp. Isolated from Lake Taihu. International Journal of Environmental Research and Public Health. 2017; 14(10):1187. https://0-doi-org.brum.beds.ac.uk/10.3390/ijerph14101187

Chicago/Turabian Style

Zhang, Juan, Qingqing Lu, Qin Ding, Lihong Yin, and Yuepu Pu. 2017. "A Novel and Native Microcystin-Degrading Bacterium of Sphingopyxis sp. Isolated from Lake Taihu" International Journal of Environmental Research and Public Health 14, no. 10: 1187. https://0-doi-org.brum.beds.ac.uk/10.3390/ijerph14101187

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop