Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Environmental Research 208 (2022) 112738

Contents lists available at ScienceDirect

Environmental Research
journal homepage: www.elsevier.com/locate/envres

A quantitative, high-throughput urease activity assay for comparison and


rapid screening of ureolytic bacteria
Ming-Juan Cui a, b, 1, Aloysius Teng c, 1, Jian Chu b, **, Bin Cao b, c, *
a
College of Civil Engineering, Fuzhou University, Fuzhou, 350108, Fujian, China
b
School of Civil and Environmental Engineering, Nanyang Technological University, 50 Nanyang Ave, Singapore, 639798, Singapore
c
Singapore Centre for Environmental Life Sciences Engineering, Interdisciplinary Graduate Programme, Graduate College, Nanyang Technological University, 60
Nanyang Dr, Singapore, 637551, Singapore

A R T I C L E I N F O A B S T R A C T

Keywords: Urease is a dinickel enzyme commonly found in numerous organisms that catalyses the hydrolysis of urea into
Urease ammonia and carbon dioxide. The microbially induced carbonate precipitation (MICP) process mediated by
Urease-producing bacteria urease-producing bacteria (UPB) can be used for many applications including, environmental bioremediation,
MICP
soil improvement, healing of cracks in concrete, and sealing of rock joints. Despite the importance of urease and
Biomineralization
UPB in various applications, a quantitative, high-throughput assay for the comparison of urease activity in UPB
and rapid screening of UPB from diverse environments is lacking. Herein, we reported a quantitative, 96-well
plate assay for urease activity based on the Christensen’s urea agar test. Using this assay, we compared urease
activity of six bacterial strains (E. coli BL21, P. putida KT2440, P. aeruginosa PAO1, S. oneidensis MR-1, S. pasteurii
DSM 33, and B. megaterium DSM 319) and showed that S. pasteurii DSM 33 exhibited the highest urease activity.
We then applied this assay to quantify the inhibitory effect of calcium on urease activity of S. pasteurii DSM 33.
No significant inhibition was observed in the presence of calcium at concentrations below 10 mM, while the
urease activity decreased rapidly at higher concentrations. At a concentration higher than 200 mM, calcium
completely inhibited urease activity under the tested conditions. We further applied this assay to screen for
highly active UPB from a wastewater enrichment and identified a strain of S. pasteurii exhibiting a substantially
higher urease activity than DSM 33. Taken together, we established a 96-well plate-based quantitative, high-
throughput urease activity assay that can be used for comparison and rapid screening of UPB. As UPB and
urease activity are of interest to environmental, civil, and medical researchers and practitioners, we envisage
wide applications of the assay reported in this study.

1. Introduction Mian and Qian, 2016), and sealing of rock joints underground (Wu et al.,
2019). In particular, UPB-mediated MICP provides an efficient and
Urease (urea amidohydrolase; EC 3.5.1.5) is a dinickel enzyme that cost-effective biomineralization approach for environmental applica­
catalyses the hydrolysis of urea to ammonia and carbon dioxide. It is the tions. For example, MICP can be used to immobilize subsurface con­
first identified metalloenzyme commonly found in bacteria, fungi, and taminants such as toxic radionuclides (Fujita et al., 2008; Lauchnor
plants (Mobley and Hausinger, 1989; Mobley et al., 1995; Krajewska and et al., 2013) and heavy metals (Cuaxinque-Flores et al., 2020; Fang et al.,
Ureases, 2009). Urease-producing bacteria (UPB), i.e., ureolytic bacte­ 2021; Yong et al., 2021), seal fluid leakages from underground pipes
ria, are capable of inducing precipitation of calcite through the release of (Phillips et al., 2016), and treat hypersaline water generated during the
carbonate ions and the increase in pH. Microbially induced calcite production of shale oil and gas (Lei et al., 2021). In addition, MICP can
precipitation (MICP) enabled by UPB has been harnessed for many ap­ also be used to mitigate the impact of climate change as it can enhance
plications, including soil improvement (Ivanov and Chu, 2008; van the capture and sequestration of atmospheric carbon dioxide into un­
Paassen et al., 2010), healing of cracks in concrete (Jonkers et al., 2010; derground geological formations (Mitchell et al., 2010; Okyay et al.,

* Corresponding author. School of Civil and Environmental Engineering, Nanyang Technological University, 50 Nanyang Ave, Singapore, 639798, Singapore.
** Corresponding author.
E-mail addresses: cjchu@ntu.edu.sg (J. Chu), bincao@ntu.edu.sg (B. Cao).
1
These authors contributed equally to this work.

https://doi.org/10.1016/j.envres.2022.112738
Received 30 November 2021; Received in revised form 12 January 2022; Accepted 12 January 2022
Available online 15 January 2022
0013-9351/© 2022 Elsevier Inc. All rights reserved.
M.-J. Cui et al. Environmental Research 208 (2022) 112738

2016) and lessen the impact of thermal stress on coral reefs (Biscéré validity of our proposed assay, we quantitatively compared urease ac­
et al., 2018). tivity of selected bacteria, including two well-known ureolytic bacteria,
Because of the ubiquitous presence and wide applications of UPB, Sporosarcina pasteurii and Bacillus megaterium. We then applied this assay
several methods have been developed to assess urease activity (Taba­ to quantify the inhibitory effect of calcium on urease activity of
tabai, 1994; Tabatabai et al., 2002). For example, Stuart’s urea broth can S. pasteurii. Finally, we demonstrated the feasibility of using this assay to
be used to screen for the presence of ureolytic activity in environmental screen for highly active ureolytic bacteria from environmental samples.
and clinical samples. This assay is based on the colour change of the
phenol red indicator caused by alkalinisation of the media during urea 2. Materials and methods
hydrolysis (Brink, 2010). It is often used to qualitatively differentiate
organisms with high urease activity such as Proteus from other genera 2.1. Bacterial strains and culture conditions
within the Enterobacteriaceae. The Christensen’s urea agar test is another
method for screening UPB by observing a similar colour change of the Bacterial strains used in this study were Escherichia coli BL21, Pseu­
phenol red indicator in the agar. Compared with the Stuart’s urea broth, domonas putida KT2440, Pseudomonas aeruginosa PAO1, Shewanella
the Christensen’s urea agar is supplemented with peptone and glucose so oneidensis MR-1, S. pasteurii DSM 33, and B. megaterium DSM 319.
that it can be used to screen a higher diversity of bacteria, including S. pasteurii DSM 33 and B. megaterium DSM 319 were purchased from the
those that are unable to use ammonia as the sole nitrogen source. German Collection of Microorganisms and Cell Cultures GmbH, while
Furthermore, the reduced buffer capacity of the Christensen’s urea agar the other strains were obtained from our current laboratory stocks. All
enables more sensitive detection of ammonia, allowing the screening of strains were maintained in 25% glycerol stock stored at − 80 ◦ C and
UPB that fail to produce sufficient ammonia to be detected in Stuart’s grown in their respective liquid or agar media when required.
urea broth (Brink, 2010). Both the Stuart’s and Christensen’s B. megaterium DSM 319 was grown in Luria-Bertani (LB) broth supple­
biochemical tests for urease activity are qualitative in nature. Urease mented with 0.1 mM NiCl2 while S. pasteurii DSM 33 was grown in an
activity can be quantified by measuring the release of ammonia using ammonium-yeast extract medium containing 20 g/L yeast extract, 15 g/
the Nessler method (Krug et al., 1979). However, this method requires L ammonium chloride and 0.1 mM NiCl2. The ammonium-yeast extract
the use of toxic mercuric salts, and the preparation of reagents is medium was then adjusted to a pH of 9.25 using NaOH. All the other
time-consuming. Another method quantifies ammonia using the Ber­ strains were grown in LB Miller broth at either 30 or 37 ◦ C.
thelot reagent (Cussac et al., 1992; van Vliet et al., 2001), which requires
sonication and incubation for a prolonged period of time. Although
several high-throughput methods for the screening and assay of urease 2.2. 96-well plate-based urease activity assay
activity have been proposed, they either require long periods of incu­
bation (Onal Okyay and Frigi Rodrigues, 2013), the use of heavy inoc­ A 96-well plate-based urease activity assay was developed using the
ulum (Qadri et al., 1984a, 1984b) or consumption of large volumes of Christensen’s urea agar (Fig. 1). The Christensen’s urea agar comprised
reagents (Cordero et al., 2019). 1 g/L of peptone, 1 g/L of dextrose, 5 g/L of sodium chloride, 2 g/L of
The objective of this study was to develop a quantitative, high- potassium phosphate, 20 g/L of urea, 0.012 g/L of phenol red, and 5 g/L
throughput urease activity assay for rapid assessment of bacterial ure­ of agar. Bacterial cells were grown in 24-well plates and then harvested
olytic potential. Specifically, we established a quantitative, 96-well plate by centrifugation and resuspended in PBS buffer. 200 μL of the cell
assay based on the Christensen’s urea agar test. To demonstrate the suspension was introduced into each well of the 96-well plate followed
by pelleting the cells by centrifugation and removal of supernatant. Cells

Fig. 1. Schematic illustration of the 96-well plate-


based urease activity assay. Tested bacteria were
grown in 24-well plates before the cells were pelleted
and washed with PBS. Then, 200 μL of cell suspension
was transferred into each well of the 96-well plate.
After the cells have been pelleted to the bottom, 200
μL of the Christensen’s urea agar was added into each
well. A layer of sterile cotton gauze moistened with
hydrochloric acid was used to cover the plate to
prevent cross-well contamination by ammonia gas
released by urease-active bacteria. The absorbance at
560 nm was measured for each well to quantify ure­
ase activity.

2
M.-J. Cui et al. Environmental Research 208 (2022) 112738

were subsequently covered by 200 μL of the Christensen’s urea agar. A X software according to the protocol described by Hall (2013). The
thin layer of sterile cotton gauze soaked with hydrochloric acid was used phylogenetic tree was constructed using the maximum-likelihood algo­
to cover the 96-well plate to prevent cross-well contamination by rithm, and bootstrap values were obtained from an analysis of 1000
gaseous ammonia. The colour change of the agar with respect to time re-samplings of the alignments.
was then monitored by measuring the absorbance at 560 nm using a
Tecan Infinite Pro II microplate reader. 3. Results and discussion

2.3. Validation of the 96-well plate-based urease activity assay 3.1. Quantitative comparison of urease activity

To validate the 96-well plate-based urease activity assay, we tested The relative urease activity of 6 bacterial strains quantified using the
urease-positive bacteria S. pasteurii DSM 33 and B. megaterium DSM 319, 96-well plate-based urease activity assay is shown in Fig. 2. S. pasteurii
urease-negative bacteria E. coli BL21 and S. oneidensis MR-1, as well as DSM 33 exhibited the fastest colour change. The rapid change of colour
P. aeruginosa PAO1 and P. putida KT2440. Fresh colonies of each bac­ indicated a high urease activity of S. pasteurii DSM 33. For E. coli BL21
terial strain were picked and inoculated into their respective growth and S. oneidensis MR-1, no change in OD560 was observed within 3 h,
media in 24-well plates and incubated at 30 ◦ C for approximately 5 h. corroborating the lack of urease activity in these two strains.
The cells for urease activity assay should be harvested at the exponential Interestingly, although B. megaterium DSM 319 is known to be
phase. All the tested cultures reached the exponential growth phase after urease-positive, no apparent change in OD560 in the first 3 h was
5-h incubation. The cell cultures were then washed three times with 1.5 observed. Similarly, neither P. aeruginosa PAO1 nor P. putida KT2440
mL of phosphate-buffered saline (PBS) and spun down at 4000 rpm for 2 showed apparent change in OD560 despite being shown capable of pro­
min. The bacterial cells were then adjusted to the same OD600 (~ 0.10) ducing urease. This could be attributed to the lower urease activity of
before 200 μL of bacterial cell culture was transferred into the 96-well these strains than S. pasteurii DSM 33 that would require a longer re­
plate in triplicates. The cells were then pelleted to the bottom of the action time for them to release the necessary amount of ammonia to
wells and 200 μL of the Christensen’s urea agar was added into each overcome the buffer capacity of the Christensen’s urea agar. In addition,
well. the nutrients present in urease agar used to maintain bacteria growth
and urease activity might be another factor contributing to the lower
2.4. Quantification of inhibitory effect of calcium on urease activity urease activity (Stewart, 1965).

S. pasteurii DSM 33 was used as a model urease-positive organism to 3.2. Quantification of inhibitory effect of calcium on urease activity
quantify inhibition of calcium on urease activity. This organism was
chosen because of its high urease activity and wide applications in By using S. pasteurii DSM 33 as a model ureolytic organism, the
geotechnical and environmental engineering (Ivanov and Chu, 2008; inhibitory effect of calcium on urease activity was quantified using the
Phillips et al., 2016; Mitchell et al., 2010). Filter-sterilized calcium 96-well plate-based urease activity assay. No significant difference was
chloride stock solution was diluted with PBS and mixed with the observed in the relative activity of urease in the presence of calcium at
Christensen’s urea agar to achieve different final Ca2+ concentrations (5, concentrations below 10 mM, while the relative activity decreased
10, 25, 50, 75, 100, 200, 250, 500 mM) and then transferred into the rapidly at higher calcium concentrations (Fig. 3 and Figure S1). The
96-well plate. Cells from an overnight culture of S. pasteurii DSM 33 were inhibition of calcium on urease activity followed the non-competitive
harvested and washed three times with PBS and the cell pellets were inhibition model (R2 = 0.99).
then resuspended in 5 mL of PBS before 200 μL of the cell suspension was Calcite precipitation induced by UPB has many applications such as
transferred into each well. The absorbance at 560 nm was measured in improving the strength of soil, healing concrete cracks and plugging of
15 min intervals. Relative urease activity at various Ca2+ concentrations rock joints (Ivanov and Chu, 2008; Jonkers et al., 2010; Wu et al., 2019).
was indicated by the rate of increase in OD560 from t = 15 min to t = 30 One key factor influencing the efficiency of the mineralization process
min. Model fitting of the relative activity at different calcium concen­ and strength of the improved soil is the amount of calcium carbonate
trations was carried out using the DoseResp function of OriginLab Pro. produced in each treatment which in turn relies on the calcium content
The non-competitive inhibition model was used: present in the cementation solution used for MICP (Krajewska, 2018;
Tang et al., 2020). Several studies have shown that a high concentration
v = Vm (
1
)(
[S]
) of calcium allows more rapid formation of calcium carbonate crystals
1 + K[I]I KM + [S] responsible for the strengthening of soil (Qabany et al., 2012; Ng et al.,
2014). However, excessive levels of calcium in the cementation solution
where v is the rate of urea hydrolysis, Vm is the maximum rate of urea will inhibit the urease activity and is detrimental to MICP (Nemati and
hydrolysis, [S] is the concentration of urea, [I] is the concentration of Voordouw, 2003; Nemati et al., 2005; Lai et al., 2021). Therefore, one
Ca2+, KI is the inhibition constant, and KM is the Michaelis constant. key factor to improve the efficiency of any MICP based methods is to
obtain a urease enzyme that can tolerate a higher calcium concentration.
2.5. Screening for ureolytic bacteria from wastewater Tools that can quantify calcium tolerance of UPB in a high-throughput
manner would facilitate the search for highly tolerant bacterial
Rapid screening for ureolytic bacteria from a wastewater enrichment strains. The 96-well plate-based urease activity assay developed in this
(Yang et al., 2020) was carried out using the 96-well plate-based urease study was demonstrated to have the capability of quantifying the
activity assay. Serial dilutions of an overnight culture of the enrichment inhibitory effect of calcium on urease activity in UPB. The high R2 value
were plated onto the ammonium-yeast extract agar. After an overnight of the fitting curve obtained shows a good fit between the experimental
incubation at 30 ◦ C, individual colonies with unique morphologies were OD560 absorbance values with a non-competitive inhibition model. The
picked for the 96-well plate-based screening of urease activity. Colony assay developed here may also be employed to quantify the inhibitory
PCR was carried out to amplify the 16S rRNA gene with the primers 27F effects of other chemicals or metals on urease (Shaw, 1954).
(5′ -AGAGTTTGATCMTGGCTCAG-3′ ) and 1492R (5′ -TACGGYTACCTT
GTTACGACTT-3′ ) using Q5 High-Fidelity DNA polymerase (NEB, USA). 3.3. Screening for ureolytic bacteria from wastewater
Clean up of the PCR products was carried out using the Wizard SV Gel
and PCR Clean-Up System (Promega, USA) before they were sent for The 96-well plate-based urease activity assay was employed to
sequencing. Phylogenetic analysis was then carried out using the MEGA screen for bacteria with high urease activity from a previously reported

3
M.-J. Cui et al. Environmental Research 208 (2022) 112738

Fig. 2. Christensen’s urea agar assay of common laboratory strains. a. Colour change of Christensen’s urea agar for the laboratory strains tested from 0 to 3 h. b.
OD560 readings of the Christensen’s urea agar with respect to time. c. Relative urease activity of the tested bacterial strains with respect to S. pasteurii DSM 33. (For
interpretation of the references to colour in this figure legend, the reader is referred to the Web version of this article.)

Fig. 3. The inhibitory effect of calcium on urease


activity in S. pasteurii DSM 33 quantified using the 96-
well plate-based assay. a. OD560 in the presence of
CaCl2. b. Relative urease activity at different calcium
concentrations. Relative urease activity was deter­
mined as the rate of colour change (OD560) of the
Christensen’s urea agar from t = 15 min to t = 30
min. Fitting curve was generated by fitting experi­
mental data to the non-competitive inhibition model
(R2 = 0.99). (For interpretation of the references to
colour in this figure legend, the reader is referred to
the Web version of this article.)

(Yang et al., 2020) wastewater enrichment. Four strains A, B, C, and D highest 16S rRNA gene similarity with S. pasteurii although the urease
were isolated from the enriched culture based on distinct colony mor­ activity assay showed that strain B had a much higher urease activity
phologies. Urease activity of the 4 isolates and the model urease-positive than S. pasteurii DSM 33. The difference in the ureolytic potential of
bacterium S. pasteurii DSM 33 was compared using the 96-well different bacterial strains from the same species could be attributed to
plate-based assay. Intriguingly, strain B exhibited higher rate of in­ strain-specific differences in their urease-encoding genes (Hammes
crease in OD560 than all the other strains, including S. pasteurii DSM 33 et al., 2003). Further genomic and transcriptomic analyses are required
(Fig. 4 and Figure S2). Both strains A and C showed a slightly higher rate in the future to understand strain-specific differences in urease activity.
of increase in OD560 than S. pasteurii DSM 33 while no increase in OD560
for strain D was observed (Fig. 4 and Figure S2). These results suggested 4. Conclusions
that among all the tested strains, strain B has the highest urease activity.
Phylogenetic analyses suggested that strains A and B showed a high In this paper, we reported a 96-well plate-based quantitative, high-
degree of sequence homology to S. pasteurii. Strain C is most closely throughput urease activity assay. Using this assay, we quantitatively
related to S. pasteurii BNCC 337394, while strain D is closely related to compared urease activity of six bacterial strains (E. coli BL21, P. putida
the uncultured Sporosarcina sp. clone FACULANA16. Detailed analysis of KT2440, P. aeruginosa PAO1, S. oneidensis MR-1, S. pasteurii DSM 33, and
the full genome of strains C and D is required to better understand their B. megaterium DSM 319) and showed that S. pasteurii DSM 33 exhibited
phylogenetic ancestry due to the low bootstrap value obtained. Strains the highest urease activity, while urease activity in E. coli BL21 and
A, B and C are likely to be aerobic, urease-positive and Gram-positive S. oneidensis MR-1 was undetectable. We then applied this assay to
bacteria from the genus of Sporosarcina or Bacillus (Claus et al., 2006). quantify the inhibitory effect of calcium on urease activity of S. pasteurii
Interestingly, among the three strains analysed, strain B showed the DSM 33. No significant inhibition was observed in the presence of

4
M.-J. Cui et al. Environmental Research 208 (2022) 112738

Fig. 4. a. Urease activity assay of the isolates A, B, C and D from a wastewater enrichment. DSM 33 and E. coli BL21 were included for comparison. b. Relative urease
activity of the isolates with reference to S. pasteurii DSM 33 and E. coli BL21. c. Phylogenetic analysis of the isolates A, B, C and D from the wastewater enrichment.
The phylogenetic tree was constructed based on the 16S rRNA sequencing results obtained from the unique colonies and other closely related bacteria, including
those that has not been cultured before. Bootstrap values obtained from 1000 re-samplings of the data are represented on the nodes.

calcium at concentrations below 10 mM, while the relative activity of provided by the Ministry of Education (MOE), Singapore, the Centre for
urease decreased rapidly at higher concentrations. At a concentration Urban Solutions, Nanyang Technological University, Singapore, and the
higher than 200 mM, calcium completely inhibited urease activity in National Research Foundation and MOE Singapore, under its Research
S. pasteurii DSM 33 under the tested conditions. We further applied our Centre of Excellence Programme, Singapore Centre for Environmental
proposed assay to screen for highly active UPB from a wastewater Life Sciences Engineering (SCELSE) (M4330005.C70 to B.C.), Nanyang
enrichment and identified a strain of S. pasteurii (strain B) exhibiting a Technological University, Singapore.
substantially higher urease activity than DSM 33. Taken together, we
established a 96-well plate-based quantitative, high-throughput urease Appendix A. Supplementary data
activity assay that can be used for comparison and rapid screening of
ureolytic bacteria. As ureolytic bacteria and urease activity are of in­ Supplementary data to this article can be found online at https://doi.
terest to environmental, civil, and medical researchers and practitioners, org/10.1016/j.envres.2022.112738.
we envisage wide applications of the assay reported in this study.
References
Declaration of competing interest
Biscéré, T., Ferrier-Pagès, C., Grover, R., Gilbert, A., Rottier, C., Wright, A., Payri, C.,
The authors declare that they have no known competing financial Houlbrèque, F., 2018. Enhancement of coral calcification via the interplay of nickel
and urease. Aquat. Toxicol. 200, 247–256. https://doi.org/10.1016/j.
interests or personal relationships that could have appeared to influence
aquatox.2018.05.013.
the work reported in this paper. Brink, B., 2010. Urease test protocol. Washington DC: Am. Soc. Microbiol.
Claus, D., Fritze, D., Kocur, M., 2006. Genera related to the genus
Acknowledgements Bacillus—sporolactobacillus, Sporosarcina, planococcus, filibacter and caryophanon.
In: Dworkin, M., Falkow, S., Rosenberg, E., Schleifer, K.-H., Stackebrandt, E. (Eds.),
The Prokaryotes: Volume 4: Bacteria: Firmicutes, Cyanobacteria. Springer US, New
This research was supported by Grant No MOE2015-T2-2-142 York, NY. https://doi.org/10.1007/0-387-30744-3.

5
M.-J. Cui et al. Environmental Research 208 (2022) 112738

Cordero, I., Snell, H., Bardgett, R.D., 2019. High throughput method for measuring Nemati, M., Voordouw, G., 2003. Modification of porous media permeability, using
urease activity in soil. Soil Biol. Biochem. 134, 72–77. https://doi.org/10.1016/j. calcium carbonate produced enzymatically in situ. Enzym. Microb. Technol. 33 (5),
soilbio.2019.03.014. 635–642. https://doi.org/10.1016/S0141-0229(03)00191-1.
Cuaxinque-Flores, G., Aguirre-Noyola, J.L., Hernández-Flores, G., Martínez-Romero, E., Nemati, M., Greene, E.A., Voordouw, G., 2005. Permeability profile modification using
Romero-Ramírez, Y., Talavera-Mendoza, O., 2020. Bioimmobilization of toxic metals bacterially formed calcium carbonate: comparison with enzymic option. Process
by precipitation of carbonates using Sporosarcina luteola: an in vitro study and Biochem. 40 (2), 925–933. https://doi.org/10.1016/j.procbio.2004.02.019.
application to sulfide-bearing tailings. Sci. Total Environ. 724, 138124. https://doi. Ng, W.S., Lee, M.L., Tan, C.K., Hii, S.L., 2014. Factors affecting improvement in
org/10.1016/j.scitotenv.2020.138124. engineering properties of residual soil through microbial-induced calcite
Cussac, V., Ferrero, R.L., Labigne, A., 1992. Expression of Helicobacter pylori urease precipitation. J. Geotech. Geoenviron. Eng. 140 (5), 04014006 https://doi.org/
genes in Escherichia coli grown under nitrogen-limiting conditions. J. Bacteriol. 174 10.1061/(ASCE)GT.1943-5606.0001089.
(8), 2466–2473. https://doi.org/10.1128/JB.174.8.2466-2473.1992. Okyay, T.O., Nguyen, H.N., Castro, S.L., Rodrigues, D.F., 2016. CO2 sequestration by
Fang, L., Niu, Q., Cheng, L., Jiang, J., Yu, Y.Y., Chu, J., Achal, V., You, T., 2021. Ca- ureolytic microbial consortia through microbially-induced calcite precipitation. Sci.
mediated alleviation of Cd2+ induced toxicity and improved Cd2+ Total Environ. 572, 671–680. https://doi.org/10.1016/j.scitotenv.2016.06.199.
biomineralization by Sporosarcina pasteurii. Sci. Total Environ. 787, 147627. Onal Okyay, T., Frigi Rodrigues, D., 2013. High throughput colorimetric assay for rapid
https://doi.org/10.1016/j.scitotenv.2021.147627. urease activity quantification. J. Microbiol. Methods 95 (3), 324–326. https://doi.
Fujita, Y., Taylor, J.L., Gresham, T.L.T., Delwiche, M.E., Colwell, F.S., McLing, T.L., org/10.1016/j.mimet.2013.09.018.
Petzke, L.M., Smith, R.W., 2008. Stimulation of microbial urea hydrolysis in Phillips, A.J., Cunningham, A.B., Gerlach, R., Hiebert, R., Hwang, C., Lomans, B.P.,
groundwater to enhance calcite precipitation. Environ. Sci. Technol. 42 (8), Westrich, J., Mantilla, C., Kirksey, J., Esposito, R., Spangler, L., 2016. Fracture
3025–3032. https://doi.org/10.1021/es702643g. sealing with microbially-induced calcium carbonate precipitation: a field study.
Hall, B.G., 2013. Building phylogenetic trees from molecular data with MEGA. Mol. Biol. Environ. Sci. Technol. 50 (7), 4111–4117. https://doi.org/10.1021/acs.est.5b05559.
Evol. 30 (5), 1229–1235. https://doi.org/10.1093/molbev/mst012. Qabany, A.A., Soga, K., Santamarina, C., 2012. Factors affecting efficiency of microbially
Hammes, F., Boon, N., de Villiers, J., Verstraete, W., Siciliano, S.D., 2003. Strain-specific induced calcite precipitation. J. Geotech. Geoenviron. Eng. 138 (8), 992–1001.
ureolytic microbial calcium carbonate precipitation. Appl. Environ. Microbiol. 69 https://doi.org/10.1061/(ASCE)GT.1943-5606.0000666.
(8), 4901–4909. https://doi.org/10.1128/AEM.69.8.4901-4909.2003. Qadri, S.M.H., Zubairi, S., Hawley, H.P., Ramirez, E.G., 1984a. Simple spot test for rapid
Ivanov, V., Chu, J., 2008. Applications of Microorganisms to geotechnical engineering detection of urease activity. J. Clin. Microbiol. 20 (6), 1198–1199. https://doi.org/
for bioclogging and biocementation of soil in situ. Rev. Environ. Sci. Biotechnol. 7 10.1128/JCM.20.6.1198-1199.1984.
(2), 139–153. https://doi.org/10.1007/s11157-007-9126-3. Qadri, S.M.H., Zubairi, S., Hawley, H.P., Mazlaghani, H.H., Ramirez, E.G., 1984b. Rapid
Jonkers, H.M., Thijssen, A., Muyzer, G., Copuroglu, O., Schlangen, E., 2010. Application test for determination of urea hydrolysis. Antonie Leeuwenhoek 50 (4), 417–423.
of bacteria as self-healing agent for the development of sustainable concrete. Ecol. https://doi.org/10.1007/BF00394656.
Eng. 36 (2), 230–235. https://doi.org/10.1016/j.ecoleng.2008.12.036. Shaw, W.H.R., 1954. The inhibition of urease by various metal ions. J. Am. Chem. Soc.
Krajewska, B., 2018. Urease-aided calcium carbonate mineralization for engineering 76 (8), 2160–2163. https://doi.org/10.1021/ja01637a034.
applications: a review. J. Adv. Res. 13, 59–67. https://doi.org/10.1016/j. Stewart, D.J., 1965. The urease activity of fluorescent pseudomonads. J. Gen. Microbiol.
jare.2017.10.009. 41 (2), 169–174. https://doi.org/10.1099/00221287-41-2-169.
Krajewska, B., Ureases, I., 2009. Functional, catalytic and kinetic properties: a review. Tabatabai, M.A., 1994. Soil enzymes. In: Methods of Soil Analysis. John Wiley & Sons,
J. Mol. Catal. B Enzym. 59 (1–3), 9–21. https://doi.org/10.1016/j. Ltd, pp. 775–833. https://doi.org/10.2136/sssabookser5.2.c37.
molcatb.2009.01.003. Tabatabai, M.A., Dick, W., 2002. Enzymes in soil: research and developments in
Krug, F.J., Růžička, J., Hansen, E.H., 1979. Determination of ammonia in low measuring activities. In: Burns, R., Dick, R. (Eds.), Enzymes in the Environment, vol.
concentrations with nessler’s reagent by flow injection analysis. Analyst 104 (1234), 84. CRC Press. https://doi.org/10.1201/9780203904039.ch21. Books in Soils,
47. https://doi.org/10.1039/an9790400047. Plants, and the Environment.
Lai, H.-J., Cui, M.-J., Wu, S.-F., Yang, Y., Chu, J., 2021. Retarding effect of concentration Tang, C.-S., Yin, L., Jiang, N., Cheng, Z., Hao, Z., Hao, L., Bin, S., 2020. Factors affecting
of cementation solution on biocementation of soil. Acta Geotech 16 (5), 1457–1472. the performance of microbial-induced carbonate precipitation (MICP) treated soil: a
https://doi.org/10.1007/s11440-021-01149-1. review. Environ. Earth Sci. 79 (5), 94. https://doi.org/10.1007/s12665-020-8840-9.
Lauchnor, E.G., Schultz, L.N., Bugni, S., Mitchell, A.C., Cunningham, A.B., Gerlach, R., van Paassen, L.A., Ghose, R., van der Linden, T.J.M., van der Star, W.R.L., van
2013. Bacterially induced calcium carbonate precipitation and strontium Loosdrecht, M.C.M., 2010. Quantifying biomediated ground improvement by
coprecipitation in a porous media flow System. Environ. Sci. Technol. 47 (3), ureolysis: large-scale biogrout experiment. J. Geotech. Geoenviron. Eng. 136 (12),
1557–1564. https://doi.org/10.1021/es304240y. 1721–1728. https://doi.org/10.1061/(ASCE)GT.1943-5606.0000382.
Lei, H., Wang, H., Pei, X., Zhang, Y., 2021. Biomineralization of hypersaline produced van Vliet, A.H.M., Kuipers, E.J., Waidner, B., Davies, B.J., de Vries, N., Penn, C.W.,
water using microbially induced calcite precipitation. Water Res. 190, 116753. Vandenbroucke-Grauls, C.M.J.E., Kist, M., Bereswill, S., Kusters, J.G., 2001. Nickel-
https://doi.org/10.1016/j.watres.2020.116753. responsive induction of urease expression in Helicobacter pylori is mediated at the
Mian, L., Qian, C., 2016. Influences of bacteria-based self-healing agents on cementitious transcriptional level. Infect. Immun. 69 (8), 4891–4897. https://doi.org/10.1128/
materials hydration kinetics and compressive strength. Construct. Build. Mater. 121, IAI.69.8.4891-4897.2001.
659–663. https://doi.org/10.1016/j.conbuildmat.2016.06.075. Wu, C., Chu, J., Wu, S., Hong, Y., 2019. 3D characterization of microbially induced
Mitchell, A.C., Dideriksen, K., Spangler, L.H., Cunningham, A.B., Gerlach, R., 2010. carbonate precipitation in rock fracture and the resulted permeability reduction.
Microbially enhanced carbon capture and storage by mineral-trapping and Eng. Geol. 249, 23–30. https://doi.org/10.1016/j.enggeo.2018.12.017.
solubility-trapping. Environ. Sci. Technol. 44 (13), 5270–5276. https://doi.org/ Yang, Y., Chu, J., Cao, B., Liu, H., Cheng, L., 2020. Biocementation of soil using non-
10.1021/es903270w. sterile enriched urease-producing bacteria from activated sludge. J. Clean. Prod.
Mobley, H.L.T., Hausinger, R.P., 1989. Microbial ureases: significance, regulation, and 262, 121315. https://doi.org/10.1016/j.jclepro.2020.121315.
molecular characterization. Microbiol. Rev. 53, 24. Yong, Z., Chen, Z., Du, Y., Lyu, Q., Yang, Z., Yang, L., Yan, Z., 2021. Microbiologically
Mobley, H.L.T., Island, M.D., Hausinger, R.P., 1995. Molecular biology of microbial induced calcite precipitation technology for mineralizing lead and cadmium in
ureases. Microbiol. Rev. 59, 31. landfill leachate. J. Environ. Manag. 296, 113199. https://doi.org/10.1016/j.
jenvman.2021.113199.

You might also like