Uploaded by serene0212

s13205-021-02774-z

advertisement
3 Biotech
(2021) 11:218
https://doi.org/10.1007/s13205-021-02774-z
ORIGINAL ARTICLE
Identification and characterization of alkaline phosphatase gene phoX
in Microcystis aeruginosa PCC7806
Sujuan Hong1
· Qianhui Pan1 · Siyu Chen1 · Yao Zu1 · Chongxin Xu2 · Jianhong Li1
Received: 4 November 2020 / Accepted: 3 April 2021
© King Abdulaziz City for Science and Technology 2021
Abstract
PhoX is an extracellular alkaline phosphatase that is widely found in cyanobacteria and plays an important role in the conversion of extracellular organophosphorus into soluble inorganic phosphorus. However, the phoX gene has not yet been
experimentally confirmed to exist in bloom-forming Microcystis species. In this study, we identified a putative phoX gene
(GenBank accession no. ARI79942.1) in M. aeruginosa PCC7806 and overexpressed it in Escherichia coli 21 (DE3). The
expressed PhoX protein displayed phosphodiesterase and phosphomonoesterase activities. In contrast to other bacterial
PhoX proteins, which are activated mainly by C
­ a2+, Microcysits PhoX was most strongly activated by M
­ g2+, followed by
2+
2+
2+
2+
2+
­Co , ­Ca , ­Zn and M
­ n , but it was inhibited by N
­ i . Sequence analysis showed that phoX was highly conserved in the
Microcystis genus (DNA similarity > 96% between species). phoX expression responded significantly to different environmental phosphorus levels. When PCC7806 cells were cultured in phosphorus-deficient medium (BG11-P), phoX expression
reached its highest level at 2 h and then decreased to a low level at 4 h. Organophosphate induced the expression of phoX; its
expression reached the highest level at 4 h and was maintained at a high level at 6 h. Our results confirmed a putative phoX
gene and demonstrated that the phoX gene of Microcystis is conserved.
Keywords PhoX · Alkaline phosphatase · Microcystis · Phosphorus metabolism · Cyanobacterial bloom
Introduction
Cyanobacterial blooms have become commonplace in many
freshwater ecosystems (Benayache et al. 2019; Chorus and
Bartram 1999; Hudnell et al. 2008). The typical cyanobacterial genus Microcystis can proliferate rapidly in a short
time under suitable environmental conditions, forming visible blue-green oil-like films or floating foams on the water
surface (Kannan and Lenca 2012). Microcystis blooms not
only destroy hydroecosystems, but also produce microcytin
peptide toxins. Microcystis blooms are a very concerning
environmental problem worldwide (Butler et al. 2009; Duy
et al. 2000; Harke et al. 2016).
* Jianhong Li
lijianhong@njnu.edu.cn; 516389186@qq.com
1
School of Life Sciences, Nanjing Normal University,
Nanjing 210023, China
2
Laboratory for Food Quality, Safety‑State Key Laboratory
Cultivation Base of Ministry of Science and Technology,
Institute of Food Safety and Nutrition, Jiangsu Academy
of Agricultural Sciences, Nanjing 210014, China
Phosphorus (P) is an essential nutrient for Microcystis
growth and bloom occurrence (Jankowiak et al. 2019; Vuorio et al. 2020). Orthophosphate is a form of P that is directly
used by phytoplankton (Correll 1998; Lee et al. 1980; Lin
et al. 2016). However, the concentration of orthophosphate
is very low in most water bodies, and dissolved organic
phosphorus (DOP) is the main source of P for phytoplankton (Krom et al. 2005; Pulido-Villena et al. 2010). Alkaline
phosphatase (APase) plays an important role in the absorption and utilization of organophosphorus by cyanobacteria
and can convert organophosphorus into orthophosphate
(Ammerman 1991; Shun et al. 1994; Siuda and Chrost
2001).
There are three families of prokaryotic APases, PhoA,
PhoD and PhoX (Luo et al. 2009; Ragot et al. 2017), which
correspond to three organophosphorus metabolism pathways
in bacteria (White 2009). First, small molecular phospholipids are transported to the cytoplasm via uptake by the
glycerol phosphate system in the cytoplasm and are decomposed by intracellular APase PhoA and/or PhoD. Second,
the small organic phospholipid molecules enter the periplasmic space of the cell and are then broken down by APase
13
Vol.:(0123456789)
218
Page 2 of 12
in the periplasm. Third, organophosphates in the ambient
environment are hydrolysed by extracellular APase (PhoX).
PhoX proteins are distinct from the well-characterized PhoA
family (Zaheer et al. 2009) and are more widely distributed
in marine bacteria than classical PhoA proteins (Sebastian
and Ammerman 2009).
phoX is also more widely distributed algae and cyanobacteria than phoA and phoD (Ammerman 1991; Dai et al.
2016; Lin et al. 2016); however, only a few phoX genes
have been identified in these species, including genes from
Prochlorococcus, Synechococcus (Kathuria and Martiny
2011), Volvox carteri (Hallmann 1999) and Chlamydomonas
reinhardtii (Quisel et al. 1996). In the best-studied harmful cyanobacterial genus, Microcystis, although the putative
phoX gene has been studied (Harke et al. 2012; Harke and
Gobler 2013; Lin et al. 2018), it has not been experimentally verified, and little is currently known about its detailed
characteristics to date.
Based on sequence analysis, we identified a putative phoX
gene (accession no. ARI79942.1) in the Microcystis aeruginosa PCC7806 (PCC7806) genome. To reveal its function,
we expressed the gene in E. coli and determined its enzyme
properties. We also investigated the expression pattern of
phoX in PCC7806 cells in response to different P conditions
in the environment.
Materials and methods
Culture of the Microcystis strain and bacteria
Microcystis aeruginosa PCC7806 (PCC7806) was obtained
from the Pasteur Culture Collection of Cyanobacteria,
France. The strain was grown in BG11 medium (Waterbury
2006) illuminated by a bank of fluorescent lights that provided a light intensity of 3,000 lx and was shaken several
times a day at 28 ± 2 °C. E. coli DH5α, which was used for
gene cloning, and E. coli BL21 (DE3), which was used for
recombinant protein overexpression (Goldman et al. 1990),
were purchased from Novagen (Germany) and grown at
37 °C in Luria–Bertani (LB) medium supplied with appropriate antibiotics.
BLAST search for a putative phoX gene in M.
aeruginosa PCC 7806
The amino acid sequence of a verified PhoX APase (accession no. WP_138072907) from Synechococcus sp. PCC
11,901 (Kathuria and Martiny 2011) was used subjected to
a BLAST search in the GenBank database (https://​www.​
ncbi.​nlm.​nih.​gov/), and a putative phoX gene (accession no.
ARI79942.1) with 68% similarity was found in PCC7806.
13
3 Biotech
(2021) 11:218
Bioinformatics analysis of putative PhoX
The properties of the putative PhoX sequence, including its
basic physical and chemical properties, conserved domains,
transmembrane domains, signal peptides, secondary structures, N-glycosylation, O-glycosylation and phosphorylation, were analysed using the Expasy Proteomics Server
(http://​www.​expasy.​org/).
Gene expression vector construction
Two primers, phoX-F (5′-GTTT
​ AAC
​ TTT
​ AAG
​ AAG
​ GAG
​ AT​
ATC​AAC ATG​AGT​ATT​TCT​CGC​CGT​AAT​TTC​-3′) and
phoX-R (5′-AGT​GAT​GGT​GAT​GGT​GAT​GTG​TAC​ATT​
TTG​ACT​GGA​TTC​GGC​GAA-3′), were used to amplify
the DNA fragment containing the entire phoX open reading frame (accession no. ARI79942.1), and the length of
the PCR fragment was 2202 bp. The PCR products were
cloned into the pBRT7 vector to generate the pBRT7-7806
expression plasmid. The plasmid was confirmed by enzyme
digestion and gene sequencing.
The pBRT7-7806 and pBRT7 vectors were transformed
into E. coli DH5α (1:50). Transformants were screened on
plates with 100 mg/L ampicillin (Amp) and checked by
PCR. Positive colonies checked by PCR were incubated in
250 mL LB medium with 100 mg/L Amp for 8 h at 37 °C
and 250 rpm. Then, the plasmids were extracted with an
extraction kit (TIANGEN, DP103) and checked by digestion
with the endonuclease Eco32 I and sequencing.
Protein expression and purification
The recombinant plasmid pBRT7-7806 and the control
plasmid pBRT7 were transformed into E. coli BL21 (DE3).
To overexpress the protein, bacteria carrying the plasmids
were incubated in LB (with Amp) liquid medium and cultured for 3–4 h at 37 °C until the ­OD720 was approximately
0.6. Then, 0.4 mM isopropyl β-d-thiogalactoside (IPTG)
was added to the culture, which was incubated overnight
at 16 °C. The liquid cultures were centrifuged at 7000 g
for 15 min to collect the bacteria. The bacteria were disrupted by sonication in 5 mL of 50 mM phosphate buffer
with 1 mM phenylmethylsulphonyl fluoride (PMSF). The
cell lysates were centrifuged at 10,000 g for 30 min at 4 °C,
and the supernatant was used for protein purification. The
target protein was recovered with a His-Trap affinity column, and the purification protocol was provided by Smart
Life Science (Ni NTA Beads 6FF). The eluted proteins were
analysed by SDS-PAGE and Western blotting. The target
protein was immunochemically detected using a mouse antiHis antibody (GenScript, A00186). In the Western blotting
Page 3 of 12
74
91
−0.415
80,557.3
733
Theoretical pI Formula
Microcystis
aeruginosa
PCC 7806
To investigate the expression of phoX and phoA under different P supply conditions, PCC7806 cells were cultured
in BG11, P-deficient BG11 (-P) and organic-P BG11
(-P + AMP) media, and gene expression was measured at 0,
2, 4 and 6 h by qPCR.
Total RNA was extracted from the cells with an RNAprep
Plank Kit (TIANGEN, Beijing, China), and the RNA was
Molecular
weight
Expression of phoA and phoX in Microcystis
under different P conditions
Number of
amino acids
To examine the metal dependence of PhoX, apoenzymes
were prepared by dialyzing purified PhoX against elution
buffer (50 mM ­NaH2PO4, 300 mM NaCl, pH 8.0) containing 100 mM EDTA. The resulting apoenzymes were then
dialyzed against the elution buffer to remove EDTA. Apoenzyme activity was measured in the presence of 2 mM
concentrations of different divalent metal ions (­ Ca2+, ­Mn2+,
­Co2+, ­Ni2+, ­Mg2+, ­Zn2+); the control treatment lacked metal
ions.
Table 1 Predicted properties of PhoX protein in Microcystis aeruginosa PCC7806
Metalloenzyme analysis
Instability
index
To better understand the characteristics of the APase
deduced from the phoX sequence, the optimum temperature
and pH, substrate specificity and metals required for enzyme
activity were studied.
Substrate Specificity: Different possible substrates,
including glucose-1-phosphate (G-1-P), disodium 4-nitrophenyl phosphate (PNPP), adenosine monophosphate
(AMP), adenosine triphosphate (ATP), sodium pyrophosphate and glucose-6-phosphate (G-6-P), were tested by
examining inorganic phosphate (Pi) release during incubation in the presence of phosphatase.
75.87
Analysis of enzymatic characteristics
31.33 (stable)
Aliphatic index Grand average
of hydropathicity (GRAVY)
Total number
of negatively
charged residues
APase activity was measured according to the release of
paranitrophenol from paranitrophenol phosphate (pNPP)
(Roy et al. 1982). One unit of enzyme was defined as the
amount required to release 1 nmol of paranitrophenol per
min at 37 °C.
C3613H5553N
965O1106S11
Measurement of APase activity
5.23
Total number
of positively
charged residues
The N-terminal
of the sequence
considered
assay, the mouse anti-His antibody or the secondary antimouse antibody (BioWord, catalogue number AA66182)
was diluted with dilution buffer at a ratio of 1:5000 (20 mM
Tris–HCl, pH 6.8, 150 mM NaCl, 0.1% [w/v] Tween-20).
Next, the immunoblotting band signals were visualized by
enhanced chemiluminescence (ELC) (Tanon, catalogue
number 180-500), and images were obtained with a cooled
charge-coupled device (CCD) camera (Tanon-4100).
218
Met
(2021) 11:218
Organism/
Name
3 Biotech
13
218
Page 4 of 12
3 Biotech
(2021) 11:218
Fig. 1 Prediction of signal peptide in PhoX of Microcystis aeruginosa PCC7806. A Tat-way signal peptide is at N-terminal
reverse-transcribed with HiScript 1­ 32® II Q RT SuperMIX
for qPCR (Vazyme, Nanjing, China). qPCR was performed
with ­AceQ® qPCR ­SYBR® Green Master Mix (Vazyme,
Nanjing, China) on a Roche ­LightCycle® 96 real-time PCR
machine using the rnpB gene as a control (Yoshida et al.
2010).
Primers for the qPCR amplification of phoX (accession
no. ARI79942.1) and phoA (accession no. ARI81755.1)
were designed according to their respective gene sequences.
Primers for the rnpB gene were designed with reference to
the research of Yoshida et al. (2010). The gene-specific
primers used for phoX were phoX-1F (5-GAG​GGG​AAC​
CGA​CCG​AGA​-3′) and phoX-1R (5′-TGG​CAA​ATA​CCC​
AAG​CGC​-3′); those for phoA were phoA-1F (5′-CAG​GCG​
CAAC
​ AGG
​ AAA
​ GTA
​ C-3′) and phoA-1R (5′-CCAG
​ AAC
​ TT​
TGC​CAT​CTT​GCT-’); and those for rnpB were rnpB-F (5′TGC​CAC​AGA​AAA​ATA​CCG​CC-3′) and rnpB-R (5′-CTC​
CAC​CTT​GCT​CCC​CAC​-3′).
Results
Bioinformatic characteristics of the putative PhoX
of PCC7806
The putative PhoX (accession no. ARI79942.1) of
PCC7806 consisted of 735 amino residues (Table 1), and
alpha helixes were mainly distributed at its N-terminus.
13
Predictions of the posttranslational modification of the
protein indicated that it contained numerous glycosylation and phosphorylation sites. The protein was a hydrophilic protein without a transmembrane domain but
included an N-terminal Tat-way signal peptide (Fig. 1),
which conformed to the characteristics of bacterial PhoX
(Wu et al. 2007).
To compare the similarity of phoX genes in cyanobacteria, based on the PhoX protein sequence of PCC7806,
106 similar proteins were found by Blastp searches in
GenBank. A phylogenetic tree was constructed with
the MEGA program. The tree showed that Microcystis
PhoXs formed an independent branch (Fig. 2). Amino
acid sequence analysis showed that phoX was conserved
in Microcystis, and the identity was more than 98%.
However, the conservation of phoX in cyanobacteria is
low, even in unicellular cyanobacteria with close genetic
relationships. For example, the identity of PCC7806 with
Gloeocapsa sp. PCC 73,106, Synechococcus sp. PCC
7336, and Synechocystis sp. PCC6803 is only 57%, 52%
and 52%, respectively.
Expression and purification of the putative phoX
An expression vector (pBRT7-7806) carrying the putative
phoX gene and a control plasmid (pBRT7) were constructed
successfully and verified by colony PCR, plasmid digestion,
and plasmid sequencing (Fig. 3).
3 Biotech
Page 5 of 12
(2021) 11:218
218
Fig. 2 The maximum likelihood tree of alkaline phosphatase protein PhoX in Cyanobacteria. The red part is Microcystis
Figure 4 shows the results of Western blotting and SDSPAGE electrophoresis of the expressed protein. The results
showed that the His-Trap eluent contained the target protein,
which showed a size of approximately 80 kDa (Fig. 5), in
accord with the theoretical value. The measurement of APase
activities showed a value of approximately 57 U/L for the eluent with the expressed protein (Fig. 6), which was three times
higher than the value for the control, and the catalytic activity
was completely abolished by incubation at 100 °C for 15 min.
These data confirmed that the expressed protein was an APase.
Enzymatic property of the PhoX
The measurement of APase activities showed that the
optimal temperature and pH for PhoX were 37 °C and 10,
respectively (Fig. 7). The substrate specificity analysis
showed that PhoX presented phosphodiesterase and phosphomonoesterase activities (Table 2). Divalent cations were
necessary for the catalytic activity of PhoX, and the order of
catalytic activity was M
­ g2+, ­Co2+, ­Ca2+, ­Zn2+ and ­Mn2+. In
2+
contrast, ­Ni inhibited PhoX activity (Table 3).
13
218
Page 6 of 12
3 Biotech
(2021) 11:218
Fig. 3 Colony identification by
PCR and plasmid pBRT7-7806
identification by restriction. A
Colony PCR, M marker; C PCR
product of a colony. B Restriction identification, M marker;
R restriction map of plasmid
pBRT7-7806 by Eco32 I
Fig. 4 Western-bolt (up) and
SDS-PAGE (down) of expression protein of the putative
phoX gene in E. coli strain
BL21(DE3). The target protein
is about 88 KDa. Lane M:
protein marker. A Bacteria with
control plasmid (pBRT7); B
bacteria with expression plasmid (pBRT7-7806); Lane A1
and B1, total proteins of bacteria without IPTG. Lane A2 and
B2, cytoplasmic supernatant
without IPTG. Lane A3 and B3,
total proteins of bacteria with
0.4 mM IPTG. Lane A4 and B4,
cytoplasmic supernatant with
0.4 mM IPTG. Lane B4 show
the target protein PhoX is overexpressed induced by IPTG in
E. coli with pBRT7-7806
Expression of Microcystis APase genes
under different P conditions
Based on BLAST searches of alignment in NCBI, a putative
phoX gene (accession no. ARI79942.1) and a putative phoA
gene (accession no. ARI81755.1) were found in PCC7806,
but no phoD gene was found.
13
In P-sufficient BG11 medium, the expression of phoX
in PCC7806 cells was very low, while in P-deficient
BG11(-P) medium, phoX expression increased almost
ten-fold at 2 h and then decreased to a level similar to
that in BG11 cells at 4 h (Fig. 8A). phoX expression was
also induced by AMP, and its activity in AMP medium
reached the highest level at 4 h, when it was 2.5 times as
3 Biotech
Page 7 of 12
(2021) 11:218
Fig. 5 SDS-PAGE of the purified protein of the putative phoX gene
in E. coli strain BL21 (DE3). Lane M, protein marker. C, E. coli
with control plasmid (pBRT7); D, E. coli with expression plasmid
(pBRT7-7806); Lane C1 and D1, cytoplasmic supernatant. Lane C2
and D2, proteins recovered by His-tag affinity chromatography from
C1 and D1, respectively. D2 shows the target protein (square box).
218
Fig. 7 The enzyme properties of expressed PhoX. A The optimum
temperature, B optimum pH. Error bars indicate SD of three independent biological replicates
Table 2 Substrate specificity of Microcystis PhoX
Substrates
Pi liberated relative to that from ­PNPPa
PNPP
Sodium pyrophosphate
β-Glycerophosphate
ATP
AMP
G-6-P
100
261 ± 0.29
107 ± 0.24
100 ± 0.13
46 ± 0.10
39 ± 0.09
a
The Pi released from PNPP is set at 100%. All results are the mean
of three experiments. The SEMs are also provided
Fig. 6 Effect of high temperature on the enzyme activity of purified
proteins from E. coli with control (pBRT7) and expression plasmid
(pBRT7-7806). A Room temperature; B the purified proteins were
denatured at 100 ℃ for 15 min. Error bars indicate SD of three independent biological replicates. Asterisks indicate t-test significant differences (NS no significant; **, P < 0.01). APA alkaline phosphatase
activity
high as that in BG11, and a higher level was maintained
than in BG11 and BG11(-P).
The expression pattern of the putative phoA gene (intracellular APase) was very different from that of extracellular
APase phoX, and a high expression level of phoA appeared
after 4 h in BG11(-P) cells (Fig. 8B).
The total APase activity of Microcystis varied depending
on the P nutrition supply and time duration (Fig. 8C). In
p-deficient medium, total APase activity was high, while in
organophosphorus medium, APase activity was induced at
13
218
Page 8 of 12
3 Biotech
(2021) 11:218
Table 3 The ability of various divalent metal ions to restore the enzymatic activity to the apoenzyme of PhoX
Metal ion (2 mM)
APase activity restored (%)a
Ca2+
Mn2+
Co2+
Ni2+
Mg2+
Zn2+
None
221.64 ± 2.14
112.98 ± 2.25
427.70 ± 6.43
−162.77 ± 7.42
585.71 ± 1.46
123.81 ± 2.07
100.00 ± 5.86
a
Results are expressed as the percentage of activity restored when
compared with the enzyme that is not added mental ions. All results
are the means of three experiments. The SEMs are provided
4 h and then decreased to a lower level within 2 h, possibly
due to the hydrolysis of organophosphorus to compensate
for the shortage of P. These results imply that Microcystis
could quickly respond to the supply of organophosphorus
in the environment.
Discussion
PhoX in Microcystis and other microorganisms
APase plays a key role in the P metabolism of cyanobacteria
and has attracted the attention of researchers. Harke et al.
(Harke et al. 2012; Harke and Gobler 2013) found a putative
phoX gene in M. aeruginosa and showed that higher expression of the gene was induced by P stress, but the authors
did not provide the gene sequence. We searched sequences
similar to those of Harke’s phoX primers in GenBank using
BLAST and found penicillin-binding proteins in M. aeruginosa NIES843 (accession no. BAG00368.1) and PCC7806
(accession no. ARI82939.1). Disappointingly, neither of
the two genes matched the characteristics of phoX. Lin
et al. (2018) analysed 32 cyanobacterial strains according
to gene similarity (> 60%) and found that phoX-like genes
were common in cyanobacteria, including three strains of
M. aeruginosa. However, these putative phoX genes were
not confirmed by experiments. It is worth mentioning that
the PCR primers in Lin’s article did not match any phoX
gene sequence found in 19 Microcystis genomes in GenBank. Therefore, if these primers are used to measure PhoX
levels in Microcystis blooms, the levels would be seriously
underestimated. In an analysis based on the identified phoX
gene of PCC7806, we found conservation of the phoX gene
in all 19 sequenced Microcystis strains. In particular, two M.
wesenbergii strains (Fig. 2) showed very conserved phoX
genes, which have not been previously reported.
There have been some reports on phoX in bacteria and
cyanobacteria (Hallmann 1999; Kathuria and Martiny 2011;
Quisel et al. 1996; White 2009), and a putative phoX gene
13
Fig. 8 Relative expression of phoX and phoA, and total alkaline phosphatase activity. A The expression of phoX gene in medium, BG11(P), BG11, and BG11(-P + AMP). B The expression of the phoA gene
in medium, BG11(-P), BG11,and BG11(-P + AMP). C Total alkaline
phosphatase activity of Microcystis in different medium, BG11(-P),
BG11, BG11(-P + AMP). Error bars indicate SD of three independent
biological replicates. Asterisks indicate t test significant differences.
(NS no significant; *, P < 0.05; **, P < 0.01). APA alkaline phosphatase activity
3 Biotech
(2021) 11:218
Page 9 of 12
218
Fig. 9 The model of PhoX by
swissmodel. B The enlarged
version of A, in which the green
site is the C
­ a2+ binding site
Fig. 10 The model result of Microcystis PhoX
is also found in Microcystis (Harke et al. 2012; Harke and
Gobler 2013; Lin et al. 2018). However, in research on the
abundance of phoX in the environment, conserved bacterial
sequences have generally been used to design PCR primers,
which results in the omission of Microcystis phoX, and some
results have shown that there is no correlation between the
abundance of phoX and Microcystis blooms in lakes (Cao
et al. 2005; Dai et al. 2014). Our results should provide useful data for the accurate evaluation of phoX in Microcystis
blooms occurring in water.
13
218
Page 10 of 12
The specificity of PhoX in Microcystis
Studies have shown that extracellular Apases in different
species require different divalent cations. In the archaebacterium Haloarcula marismortui, the complete inhibition of extracellular APase by EDTA can be reversed by
adding ­Ca2+ ions but not by adding Z
­ n2+, ­Mn2+, or ­Mg2+
(Goldman et al. 1990). Similarly, in the green alga Volvox carteri and the azotobacter Sinorhizobium meliloti,
only ­C a 2+ can reactivate EDTA-inactivated APase,
whereas the other divalent metal cations cannot (Hallmann 1999; Zaheer et al. 2009). In Pasteurella multocida
X-73, although ­Ca2+ was shown to be the best metal ion
for restoring PhoX apoenzyme activity, C
­ o 2+ and N
­ i 2+
resulted in approximately equal activities (reactivation:
­Ca2+ 132%, C
­ o2+ 126% and N
­ i2+ 122%, respectively) (Wu
et al. 2007).
The PhoX proteins of bacteria usually contain a conserved calcium-binding domain (Wu et al. 2007; Zaheer
et al. 2009), and Microcystis PhoX also contains a nonconserved calcium-binding site (data not listed, see
Appendix), as shown by SwissModel (https://​swiss​model.​
expasy.​org). However, its apoenzyme produced the highest activity when combined with M
­ g 2+ and displayed
different levels of dependence on the other divalent
cations, which demonstrated that PhoX in Microcystis is
unique.
Phosphorus metabolism
Cyanobacteria have developed several strategies for responding to a low P supply, including compensating for their need
for P using cellular phosphates and/or synthesizing phosphate
transporters to enhance P acquisition (Orchard et al. 2009;
Tetu et al. 2009; Van Mooy et al. 2006) and increasing the
affinity and rate of P uptake (Mazard et al. 2012). The P-specific transport (Pst) system includes a periplasmic P-binding
protein (PstS) that sensitively reacts to a lack of P. The ability
to use the DOP pool, mediated by APase, is an additional
strategy for P acquisition. When P (Pitt et al. 2010) is abundant, microbes store excess P in the DOP pool as polyphosphate (poly P), and poly P is decomposed when microbes are
in a P-deficient environment (Garbisu et al. 1993).
Harke et al. (2012) and Lin et al. (2018) studied the
responses of APase genes to environmental P in Microcystis. In their studies, samples were taken every other
day, and their results showed the responses over a longer
period of time. In this study, we took samples every 2 h
to observe the more detailed responses of APase genes
to P conditions in the environment. Our results showed
that in organophosphorus medium, phoX expression was
obviously higher than that of phoA (Fig. 8A, B), which
13
3 Biotech
(2021) 11:218
implied that Microcystis increased extracellular APase
secretion to decompose organic substrates in the environment. In P-deficient medium, phoA could be induced to
decompose the poly-P substrates stored in cells. This could
be the reason that phoX expression decreased after 4 h of
P deficiency.
Accession numbers
DNA sequences: Genbank accessions ARI79942.1,
WP_138072907, ARI81755.1, BAG00368.1, and
ARI82939.1.
Appendix
See Figs. 9 and 10.
Acknowledgements The project was sponsored by The National Natural Science Foundation of China (31971561, 31370217, 31701724),
and supported by the project of Priority Academic Program Development of Jiangsu Higher Education Institutions (PAPD).
Author contributions SH and JL conceived and initiated the project.
SH, QP, SC, YZ and CX performed the experiments. SH and JL wrote
the paper. All authors read and approved the manuscript.
Declarations
Conflict of interest The authors declare that they have no conflict of
interest in the publication.
References
Ammerman JW (1991) Role of ecto-phosphohydrolases in phosphorus
regeneration in estuarine and coastal ecosystems. In: Chróst RJ
(ed) Microbial enzymes in aquatic environments. Springer, New
York, pp 165–186
Benayache NY, Nguyen-Quang T, Hushchyna K, Mclellan K, Bouacha N (2019) An overview of cyanobacteria harmful algal bloom
(CyanoHAB) issues in freshwater ecosystems. In: Gökçe D (ed)
Limnology—some new aspects of inland water ecology. IntechOpen, London, pp 1–25
Butler N, Carlisle JC, Linville R, Washburn B (2009) Microcystins: a
brief overview of their toxicity and effects, with special reference
to fish, wildlife, and livestock. OEHHA Ecotoxicol 1–2
Cao X, Ševců A, Znachor P, Zapomělová E, Liu G, Vrba J, Zhou Y
(2005) Detection of extracellular phosphatases in natural spring
phytoplankton of a shallow eutrophic lake (Donghu, China). Eur J
Phycol 40:251–258. https://d​ oi.o​ rg/1​ 0.1​ 080/0​ 96702​ 60500​ 19276​ 0
Chorus I, Bartram J (1999) Toxic cyanobacteria in water: a guide to
their public health consequences, monitoring and management. E
& FN Spon, New York
Correll D (1998) The role of phosphorus in the eutrophication of
receiving waters: a review. J Environ Qual. https://​doi.​org/​10.​
2134/​jeq19​98.​00472​42500​27000​20004x
3 Biotech
(2021) 11:218
Dai J, Chen D, Gao G, Tang X, Wu S, Wu X, Zhou J (2014) Recovery of novel alkaline phosphatase-encoding genes (phoX) from
eutrophic Lake Taihu. Can J Microbiol 60:167–171. https://​doi.​
org/​10.​1139/​cjm-​2013-​0755
Dai J, Gao G, Wu S, Wu X, Zhou J, Xue W, Yang Q (2016) Bacterial
alkaline phosphatases and affiliated encoding genes in natural
waters: a review. J Lake Sci 28:1153–1166. https://​doi.​org/​10.​
18307/​2016.​0601
Duy TN, Lam PK, Shaw GR, Connell DW (2000) Toxicology and
risk assessment of freshwater cyanobacterial (blue-green algal)
toxins in water. Rev Environ Contam Toxicol 163:113–185.
https://​doi.​org/​10.​1007/​978-1-​4757-​6429-1_3
Garbisu C, Hall DO, Serra JL (1993) Removal of phosphate by foamimmobilized Phormidium laminosum in batch and continuousflow bioreactors. Rev Environ Contam Toxicol 57:181–189.
https://​doi.​org/​10.​1002/​jctb.​28057​0214
Goldman S, Hecht K, Eisenberg H, Mevarech M (1990) Extracellular
­Ca2(+)-dependent inducible alkaline phosphatase from extremely
halophilic archaebacterium Haloarcula marismortui. J Bacteriol
172:7065–7070. https://​doi.​org/​10.​1128/​jb.​172.​12.​7065-​7070.​
1990
Hallmann A (1999) Enzymes in the extracellular matrix of Volvox: an
inducible, calcium-dependent phosphatase with a modular composition. J Biol Chem 274:1691–1697. https://​doi.​org/​10.​1074/​
jbc.​274.3.​1691
Harke MJ, Gobler CJ (2013) Global transcriptional responses of the
toxic cyanobacterium, Microcystis aeruginosa, to nitrogen stress,
phosphorus stress, and growth on organic matter. PLoS ONE
8:e69834. https://​doi.​org/​10.​1371/​journ​al.​pone.​00698​34
Harke MJ, Berry DL, Ammerman JW, Gobler CJ (2012) Molecular
response of the bloom-forming cyanobacterium, Microcystis
aeruginosa, to phosphorus limitation. Microb Ecol 63:188–198.
https://​doi.​org/​10.​1007/​s00248-​011-​9894-8
Harke MJ, Steffen MM, Gobler CJ, Otten TG, Wilhelm SW, Wood
SA, Paerl HW (2016) A review of the global ecology, genomics,
and biogeography of the toxic cyanobacterium, Microcystis spp.
Harmful Algae 54:4–20. https://d​ oi.o​ rg/1​ 0.1​ 016/j.h​ al.2​ 015.1​ 2.0​ 07
Hudnell HK, Dortch Q, Zenick H (2008) An overview of the interagency, international symposium on cyanobacterial harmful algal
blooms (ISOC-HAB): advancing the scientific understanding of
freshwater harmful algal blooms. In: Hudnell HK (ed) Cyanobacterial harmful algal blooms: state of the science and research
needs. Springer, New York, pp 1–16
Jankowiak J, Hattenrath-Lehmann T, Kramer B, Ladds M, Gobler C
(2019) Deciphering the effects of nitrogen, phosphorus, and temperature on cyanobacterial bloom intensification, diversity, and
toxicity in western Lake Erie. Limnol Oceanogr 64:1347–1370.
https://​doi.​org/​10.​1002/​lno.​11120
Kannan MS, Lenca N (2012) Field guide to algae and other "scums"
in ponds, lakes, streams and rivers. In: Press (ed) Booklet published by the Boone, Kenton and Campbell County Conservation
Districts, KY, p 5
Kathuria S, Martiny AC (2011) Prevalence of a calcium-based alkaline
phosphatase associated with the marine cyanobacterium Prochlorococcus and other ocean bacteria. Environ Microbiol 13:74–83.
https://​doi.​org/​10.​1111/j.​1462-​2920.​2010.​02310.x
Krom MD et al (2005) Summary and overview of the CYCLOPS P
addition Lagrangian experiment in the Eastern Mediterranean.
Deep-Sea Res Part II-Top Stud Oceanogr 52:3090–3108. https://​
doi.​org/​10.​1016/j.​dsr2.​2005.​08.​018
Lee G, Jones R, Rast W (1980) Availability of phosphorus to phytoplankton and its implications for phosphorus management strategies. In: Loehr RC, Martin CS, Rast W (eds) Phosphorus management strategies for lakes. Ann Arbor Sci, Ann Arbor, pp 1–42
Page 11 of 12
218
Lin SJ, Litaker RW, Sunda WG (2016) Phosphorus physiological ecology and molecular mechanisms in marine phytoplankton. J Phycol
52:10–36. https://​doi.​org/​10.​1111/​jpy.​12365
Lin WT, Zhao DD, Luo JF (2018) Distribution of alkaline phosphatase
genes in cyanobacteria and the role of alkaline phosphatase on the
acquisition of phosphorus from dissolved organic phosphorus for
cyanobacterial growth. J Appl Phycol 30:839–850. https://d​ oi.o​ rg/​
10.​1007/​s10811-​017-​1267-3
Luo HW, Benner R, Long RA, Hu JJ (2009) Subcellular localization
of marine bacterial alkaline phosphatases. Proc Natl Acad Sci
USA 106:21219–21223. https://d​ oi.o​ rg/1​ 0.1​ 073/p​ nas.0​ 90758​ 6106
Mazard S, Wilson WH, Scanlan DJ (2012) Dissecting the physiological
response to phosphorus stress in marine Synechococcus isolates
(cyanophyceae). J Phycol 48:94–105. https://​doi.​org/​10.​1111/j.​
1529-​8817.​2011.​01089.x
Orchard ED, Webb EA, Dyhrman ST (2009) Molecular analysis of the
phosphorus starvation response in Trichodesmium spp. Environ
Microbiol 11:2400–2411. https://​doi.​org/​10.​1111/j.​1462-​2920.​
2009.​01968.x
Pitt FD, Mazard S, Humphreys L, Scanlan DJ (2010) Functional characterization of Synechocystis sp. strain PCC 6803 pst1 and pst2
gene clusters reveals a novel strategy for phosphate uptake in a
freshwater cyanobacterium. J Bacteriol 192:3512–3523. https://​
doi.​org/​10.​1128/​JB.​00258-​10
Pulido-Villena E, Rerolle V, Guieu C (2010) Transient fertilizing effect
of dust in P-deficient LNLC surface ocean. Geophys Res Lett.
https://​doi.​org/​10.​1029/​2009g​l0414​15
Quisel JD, Wykoff DD, Grossman AR (1996) Biochemical characterization of the extracellular phosphatases produced by phosphorusdeprived Chlamydomonas reinhardtii. Plant Physiol 111:839–848.
https://​doi.​org/​10.​1104/​pp.​111.3.​839
Ragot SA, Kertesz MA, Meszaros E, Frossard E, Bunemann EK (2017)
Soil phoD and phoX alkaline phosphatase gene diversity responds
to multiple environmental factors. FEMS Microbiol Ecol. https://​
doi.​org/​10.​1093/​femsec/​fiw212
Roy NK, Ghosh RK, Das J (1982) Monomeric alkaline phosphatase
of Vibrio cholerae. J Bacteriol 150:1033–1039. https://d​ oi.o​ rg/1​ 0.​
1128/​JB.​150.3.​1033-​1039.​1982
Sebastian M, Ammerman JW (2009) The alkaline phosphatase PhoX
is more widely distributed in marine bacteria than the classical
PhoA. ISME J 3:563–572. https://​doi.​org/​10.​1038/​ismej.​2009.​10
Shun Y, McKelvie ID, Hart BT (1994) Determination of alkaline
phosphatase-hydrolyzable phosphorus in natural water systems
by enzymatic flow injection. Limnol Oceanogr 39:1993–2000.
https://​doi.​org/​10.​4319/​lo.​1994.​39.8.​1993
Siuda W, Chrost RJ (2001) Utilization of selected dissolved organic
phosphorus compounds by bacteria in lake water under non-limiting orthophosphate conditions. Pol J Environ Stud 10:475–483
Tetu SG, Brahamsha B, Johnson DA, Tai V, Phillippy K, Palenik B,
Paulsen IT (2009) Microarray analysis of phosphate regulation in
the marine cyanobacterium Synechococcus sp. WH8102. ISME J
3:835–849. https://​doi.​org/​10.​1038/​ismej.​2009.​31
Van Mooy BA, Rocap G, Fredricks HF, Evans CT, Devol AH (2006)
Sulfolipids dramatically decrease phosphorus demand by picocyanobacteria in oligotrophic marine environments. Proc Natl
Acad Sci U S A 103:8607–8612. https://​doi.​org/​10.​1073/​pnas.​
06005​40103
Vuorio K, Järvinen M, Kotamäki N (2020) Phosphorus thresholds for
bloom-forming cyanobacterial taxa in boreal lakes. Hydrobiologia
847:4389–4400. https://​doi.​org/​10.​1007/​s10750-​019-​04161-5
Waterbury JB (2006) The cyanobacteria—isolation, purification and
identification. In: Dworkin M, Falkow S, Rosenberg E, Schleifer
K-H, Stackebrandt E (eds) The prokaryotes. Bacteria: firmicutes,
cyanobacteria, vol 4. Springer, New York, pp 1053–1073
13
218
Page 12 of 12
White AE (2009) New insights into bacterial acquisition of phosphorus
in the surface ocean. Proc Natl Acad Sci USA 106:21013–21014.
https://​doi.​org/​10.​1073/​pnas.​09124​75107
Wu JR, Shien JH, Shieh HK, Hu CC, Gong SR, Chen LY, Chang PC
(2007) Cloning of the gene and characterization of the enzymatic
properties of the monomeric alkaline phosphatase (PhoX) from
Pasteurella multocida strain X-73. Fems Microbiol Lett 267:113–
120. https://​doi.​org/​10.​1111/j.​1574-​6968.​2006.​00542.x
Yoshida M, Yoshida T, Yoshida-Takashima Y, Kashima A, Hiroishi
S (2010) Real-time PCR detection of host-mediated cyanophage
13
3 Biotech
(2021) 11:218
gene transcripts during infection of a natural Microcystis aeruginosa population. Microbes Environ 25:211–215. https://​doi.​org/​
10.​1264/​jsme2.​me101​17
Zaheer R, Morton R, Proudfoot M, Yakunin A, Finan TM (2009)
Genetic and biochemical properties of an alkaline phosphatase
PhoX family protein found in many bacteria. Environ Microbiol
11:1572–1587. https://d​ oi.o​ rg/1​ 0.1​ 111/j.1​ 462-2​ 920.2​ 009.0​ 1885.x
Download