You are on page 1of 10

Science of the Total Environment 672 (2019) 722–731

Contents lists available at ScienceDirect

Science of the Total Environment

journal homepage: www.elsevier.com/locate/scitotenv

Short Communication

Microbial induced carbonate precipitation for immobilizing Pb


contaminants: Toxic effects on bacterial activity and
immobilization efficiency
Ning-Jun Jiang a,b,1, Rui Liu a,1, Yan-Jun Du a,⁎, Yu-Zhang Bi a
a
Jiangsu Key Laboratory of Urban Underground Engineering & Environmental Safety, Institute of Geotechnical Engineering, Southeast University, Nanjing 210096, China
b
Department of Civil and Environmental Engineering, University of Hawaii at Manoa, Honolulu, HI 96822, USA

H I G H L I G H T S G R A P H I C A L A B S T R A C T

• Pb contaminants are immobilized by


microbial induced carbonate precipita-
tion.
• Pb has toxic effects on bacterial activity
and immobilization efficiency.
• Abiotic and biotic precipitation and
bio-sorption are immobilization
mechanisms.

a r t i c l e i n f o a b s t r a c t

Article history: Microbial induced carbonate precipitation (MICP) is a natural bio-mediated process, which has been explored for
Received 22 October 2018 soil stabilization and heavy metals immobilization in soil and groundwater. Previous studies have shown that
Received in revised form 14 March 2019 MICP is capable of immobilizing various heavy metals including lead (Pb). However, most studies focus merely on
Accepted 19 March 2019
the immobilization of heavy metals with relatively low concentration. This study: (1) presents results of an investi-
Available online 20 March 2019
gation into the toxic effects of Pb on bacterial activity and immobilization efficiency within a wide range of Pb con-
Editor: Filip M.G. Tack centrations; and (2) identifies controlling biotic and abiotic factors of Pb immobilization by MICP. In the first series of
tests, bacterial strains (Sporosarcina pasteurii) are inoculated into nutrient solutions containing 0–50 mM Pb(NO3)2
Keywords: and incubated at 30 °C. Biochemical parameters are measured over time, which include pH, electrical conductivity,
Microbial induced carbonate precipitation urease activity, and viable cell number. In the second series of tests, grown bacterial strains are mixed with urea, cal-
Sporosarcina pasteurii cium salts and Pb(NO3)2 in solution. Viable cell number, produced ammonium concentration, aqueous Pb concen-
Urease activity tration of the mixed solution, and total precipitation mass are measured. The results show that the presence of Pb
Lead contamination has marginal effect on bacterial growth and associated urease activity at Pb concentration b 30 mM. The calcium
source and initial bacteria concentration are found to remarkably influence Pb immobilization efficiency in terms
of Pb removal percentage. Supplementary geochemical simulation results indicate that the Pb immobilization mech-
anisms includes abiotic precipitation, biotic precipitation and bio-sorption.
© 2019 Elsevier B.V. All rights reserved.

⁎ Corresponding author at: Institute of Geotechnical Engineering, Southeast University, Nanjing 210096, China.
E-mail addresses: jiangn@hawaii.edu (N.-J. Jiang), liurui1926@seu.edu.cn (R. Liu), duyanjun@seu.edu.cn (Y.-J. Du), biyuzhang@imde.ac.cn (Y.-Z. Bi).
1
Co-first authors. These authors contributed equally to this work.

https://doi.org/10.1016/j.scitotenv.2019.03.294
0048-9697/© 2019 Elsevier B.V. All rights reserved.
N.-J. Jiang et al. / Science of the Total Environment 672 (2019) 722–731 723

1. Introduction Although it is promising to apply MICP in heavy metal immobilization,


especially Pb, most previous studies focus on heavy metal concentration
Heavy metal contamination in soil and groundwater is a major risk no higher than 5 mM in aqueous solution or 400 mg/kg in soil (Kang
to the ecosystem and human health, since (1) heavy metals accumulate et al., 2015; Li et al., 2013; Li et al., 2016; Mugwar and Harbottle, 2016;
in air, water, soil and biosphere; and (2) industrial activities have Mwandira et al., 2017; Qian et al., 2017; Zhu et al., 2017). While these
strengthened this effect (Nagajyoti et al., 2010; Wei and Yang, 2010; studies have demonstrated the effectiveness of MICP for immobilizing ex-
Xia et al. 2019a,b). In the past few decades, various in-situ and ex-situ changeable Pb, the detrimental effects of high Pb concentration on bacte-
methods have been developed to treat heavy metal contaminated soil rial activity and MICP effectiveness have not yet been well addressed.
and groundwater. These methods include but not limited to electroki- According to the investigation by Duan et al. (2016), nearly 20% of sites
netic extraction, soil flushing, solidification/stabilization (S/S), in China are contaminated by lead, whose concentration is higher than
phytoremediation, and bioremediation. Among these remediation the regulatory limit value (80 mg/kg). A lead concentration of
methods, bioremediation utilizes microorganisms to detoxify contami- 30,430 mg/kg in soil was even detected close to a lead and zinc melting
nants via bio-sorption, extracellular chemical precipitation, and bioac- plant in Fujian Province, China. Therefore, it is of great importance to ex-
cumulation (Srivastava and Kowshik, 2018). Bioremediation has been amine the MICP immobilization efficiency at high Pb concentration.
used widely in treating wastewater and petroleum polluted soils by To conduct MICP in-situ, both bio-augmentation and bio-stimulation
stimulating native microbes to degrade contaminants, mostly toxic techniques can be used. The distinction between bio-augmentation and
organics. bio-stimulation is that bio-augmentation introduces exogenous bacteria
Microbially induced carbonate precipitation (MICP) is an emerging strains, and bio-stimulation enriches indigenous bacteria to promote
bio-mediated technique and its potential applications in civil and envi- certain biogeochemical processes (Dhami et al., 2017; Gomez et al.,
ronmental engineering have been extensively investigated. It is re- 2017). Bio-stimulation saves the cost of incubating bacteria but still
ported that MICP can be used to strengthen sandy soil, improve soil gives comparable performances in soils (Gomez et al., 2017), while
resistance to erosion, mitigate liquefaction, and seal concrete cracks bio-augmentation serves better where natural nutrients are poor
(Dejong et al., 2013). While most of these studies are conducted under (Dhami et al., 2017). In this study, we adopts the bio-augmentation
laboratory-scale conditions, a few field trials are also reported. For in- pathway for the experiment.
stance, Phillips et al. (2016) promoted MICP in the field to seal fractured This study aims to: (1) investigate the survivability of Sporosarcina
sandstone with the help of oil field technology. van Paassen (2011) ap- pasteurii and Pb immobilization efficiency by MICP within a wide
plied MICP in a field trial using injection and extraction wells to improve range of Pb concentrations; and (2) identify controlling factors that af-
the mechanical stability of horizontal boreholes. fect Pb immobilization efficiency by MICP. In this study, bacterial viabil-
MICP is initiated through the activity of enzymes such as urease ity and Pb immobilization efficiency via MICP were investigated with
(Castanier et al., 1993; Castanier et al., 1999; Gollapudi et al., 1995), car- initial Pb concentration up to 50 mM in a series of aqueous solution ex-
bonic anhydrase (Dhami et al., 2017) or asparaginase (Li et al., 2015). periments. Bacterial viability was evaluated based on electrical conduc-
One of the most popular pathways of achieving MICP is through hydro- tivity (EC), pH, viable cell numbers and urease activity at various Pb
lysis of urea by urease enzyme, which is originated from some soil bac- concentrations. Pb immobilization efficiency via MICP was assessed
teria or plants. Eqs. (1)–(4) show the biochemical reactions involving based on Pb removal percentage at various Pb concentrations, initial
ureolytic-driven MICP (van Paassen, 2009): bacteria concentrations and calcium sources.

ðNH2 Þ2 CO þ H2 O→2NH3 þ CO2 ð1Þ 2. Materials and testing methods


þ −
2NH3 þ 2H2 O↔2NH4 þ 2OH ð2Þ 2.1. Bacteria and chemicals

CO2 þ 2OH− ↔HCO3 – þ OH– ↔CO3 2– þ H2 O ð3Þ Sporosarcina pasteurii (S. pasteurii, ATCC 6452) was used in this
study for MICP. S. pasteurii is non-pathogenic bacterium with high
Ca2þ þ CO3 2– ↔CaCO3 ðsÞ ð4Þ level of urease activity. Bacterial solution was made by inoculating bac-
terial colonies into the NH4-YE solution medium (20 g/L yeast extract,
Among various MICP pathways, ureolytic-driven MICP outstands 10 g/L (NH4)2SO4, and 0.13 M Tris-base) and then subjected to
since ureolytic bacteria are widespread in natural soils (Gomez et al., shaking-incubation at 30 °C for approximately 24 h until the optical
2014) and this process is highly energy efficient in forming carbonate density at 600 nm (OD600) reached 1.5. The OD600 value was measured
precipitates (DeJong et al., 2010). So in this paper, “MICP” refers to the via a UV–Visible spectrometer (Shanghai Analytical Instruments Co Ltd.,
ureolytic-driven MICP process. While MICP is extensively studied for Shanghai, China). The prepared bacterial solution was then stored at 4
improving mechanical and hydraulic performances of granular soils °C. Prior to its use in the experiment, it was firstly re-harvested in a
and other construction materials, it is also attempted as a bioremedia- fresh growth medium. When bacteria cells reached the exponential
tion method for immobilizing heavy metal contaminants and radioac- growth stage during resuscitation, they were inoculated into the solu-
tive wastes (Fujita et al., 2004). Qian et al. (2017) used urease- tion for formal test. Stock solutions (1 M) were prepared using analyti-
producing fungi to produce carbonate precipitates and immobilize Cr cally pure reagents including Pb(NO3)2, CaCl2, Ca(Ac)2 (i.e. calcium
and Pb contained in farmland soils at 100 mg/kg and 200 mg/kg, respec- acetate) and urea. To make solutions, the predetermined amount of re-
tively. 95% exchangeable Cr and Pb in the soil were reported to be agents was mixed with proper amount of distilled water using a mag-
immobilized. Li et al. (2016) introduced another urease-producing mi- netic stirrer. The solution was then diluted with distilled water to
crobe T. tumescens capable of immobilizing multiple heavy metals (Ni, 1.0 M. NH4-YE medium, CaCl2 and calcium acetate (Ca(Ac)2) solutions
Cu Pb, Co, Zn and Cd) in aqueous solution with a removal percentage were sterilized by autoclaving at 121 °C for 20 min, while Pb(NO3)2
of approximately 90%. Fujita et al. (2010) examined the potential of na- and urea solutions were filter-sterilized with 0.22-μm filter membrane
tive ureolytic microbes for fixing 90Sr in groundwater. Achal et al. (Zhao et al., 2017). Stock solutions were stored at 4 °C before use.
(2012b) mixed extracted ureolytic bacteria from mining site with Pb-
contaminated soil (100 mg/kg) to trigger MICP and yielded a Pb re- 2.2. Bacterial activity tests
moval percentage of 83%. Co-precipitation of heavy metals in carbonate
minerals after MICP treatment was reported to be the dominant immo- Bacterial activity test was conducted to monitor the growth of newly
bilization mechanism in most of these studies. inoculated S. pasteurii in Pb-containing NH4-YE liquid medium, which
724 N.-J. Jiang et al. / Science of the Total Environment 672 (2019) 722–731

aimed to reveal the effects of Pb concentration on bacterial survivability on fresh NH4-YE agar plates. Number of bacterial colonies on the agar
and urease activity. The overview of the bacterial activity test is sche- plate was counted after 24 h incubation at 30 °C. Viable cell number in
matically shown in Fig. 1. Bacterial solution was firstly inoculated (1% final bacterial solution was expressed in CFU and was equal to counted
(v/v)) into the NH4-YE solution medium amended with Pb(NO3)2. The colony number on the plate multiplied by a dilution factor (102, 104, 106,
concentration of Pb(NO3)2 ranged from 0 to 50 mM. The final bacterial and 108 in this study) (Kang and So, 2016). All abovementioned tests
solution containing bacteria, nutrient and Pb(NO3)2 was then incubated had three biological replicates with coefficient of variance (COV) b10%.
at 30 °C with shaking (200 rpm) for 72 h. pH, EC and urease activity of
the final bacterial solution were measured at 0, 1, 2, 4, 8, 24, 48 and 2.3. Pb immobilization tests
72 h. Viable cell number in terms of colony forming unit (CFU) was
measured at 0, 12, 24, 48 and 72 h. In particular, EC measured the elec- The objective of the Pb immobilization test was to determine the Pb
trolytic ion concentrations in aqueous solution. Bacterial activity immobilization efficiency via MICP at different initial Pb and bacterial
(i.e., metabolism) involved the decomposition of large organic mole- concentrations, and calcium sources. The overview of the Pb immobili-
cules in culture medium into electrolytic ions (e.g., the consumption zation test is shown in Fig. 1. Three experimental settings were designed
of protein by bacteria), leading to increased EC in the solution to examine the effects of calcium source and initial bacteria concentra-
(Krishnamurti and Kate, 1951). Therefore, EC was used as an indicator tion on the Pb immobilization efficiency via MICP at different initial Pb
of bacterial metabolic activity in this study. concentrations. In Case 1, 3 mL bacterial solution (OD600 = 1.5) was
More specifically, pH and EC were measured using a handheld pH mixed with 27 mL aqueous solution of 0–50 mM Pb(NO3)2 and
meter (Shanghai Leici PHB-4 handheld pH meter, China) and a bench- 200 mM urea/CaCl2 mixture. In Case 2, the procedures were the same
top conductivity meter (Thermo Scientific Orion Star A215 pH/Conduc- as Case 1 except that CaCl2 was substituted by Ca(Ac)2. In Case 3, Pb
tivity benchtop multiparameter, USA), respectively. Urease activity was (NO3)2, urea and Ca(Ac)2 were dissolved in the bacterial solution to
measured based on the method presented by Van Paassen (2009) and make a total volume of 30 mL with 0–50 mM Pb(NO3)2 and urea/Ca
was expressed as the amount of hydrolyzed urea per minute. 1 mL (Ac)2 at 200 mM. It can be seen that Case 1 and Case 2 were differed
final bacterial solution was mixed with 9 mL 1.11 M urea, with EC mea- in terms of calcium source while Case 2 and Case 3 were differed in
sured at 1 min and 8 min after mixing at room temperature. Eq. (5) terms of initial bacterial concentration. The 200 mM urea/Calcium con-
gives the equation to calculate urease activity: centration was selected that did not adversely affect the viability of
S. pasteurii. It should be noted, however, that the different anions of cal-
EC8 −EC1   cium sources might impact the bacterial viability and MICP efficiency
UA ¼  10  11 mM urea min−1 ð5Þ
7 differently.
The mixed solution was incubated at 30 °C with 2 mL solution sam-
where UA is urease activity and EC1 and EC8 are electrical conductivity ples taken at 0, 24 and 48 h for the measurement of viable cell number,
at 1 and 8 min, respectively. The viable cell number was obtained produced ammonium concentration (which came from the urea hydro-
using the plate counting method (Sanders, 2012). Viable cell number lysis) and remaining Pb concentration in solution. The ammonium con-
was preferred over OD600 measurement in this study, since OD600 was centration in solution was determined using the modified Nessler
proportional to total cell number rather than viable one. Firstly, final method (Whiffin et al., 2007). The solution sample was firstly acidified
bacterial solution underwent serial dilution and was thereby spread to convert all ammonia into ammonium. Then, it was diluted 100–500

Fig. 1. Schematic diagram of the bacterial activity test (marked in red) and Pb immobilization test (marked in blue) (DW: distilled water and BS: bacterial solution). (For interpretation of
the references to colour in this figure legend, the reader is referred to the web version of this article.)
N.-J. Jiang et al. / Science of the Total Environment 672 (2019) 722–731 725

times to minimize the interference of Pb and Ca in the solution. 2 mL di- synthesizing exopolysaccharide (EPS) in response to heavy metal toxic-
luted solution was subsequently mixed with 100 μL Nessler solution for ity (Gupta and Diwan, 2017). Fig. 2(b) shows the change of solution pH
1 min before subjected to the spectrophotometer measurement at the with time in the bacterial activity test. At Pb concentration no higher
wavelength of 425 nm. Absorbance reading was then converted to the than 30 mM, pH decreased from (8.5~9.5) to (7.5–~8) in 72 h. The
ammonium concentration by referring to a calibration curve obtained drop in pH at initial hours was probably due to acid-base equilibria
using standard NH4Cl. The Pb concentration was determined using a with the addition of Pb(NO3)2, while continuous pH decrease later
Thermo Scientific iCE 3000 Series atomic absorption spectrometer was assumed to result from the production of carbonic acid by respira-
after the sample solution was acidified to yield a pH b2.0. At the end tion and organic acids by the consumption of carbohydrates (Jimenez-
of the test, produced precipitation was separated from aqueous solution Lopez et al., 2008) during bacterial metabolism. Higher Pb concentra-
using the vacuum filtration method (Keykha et al., 2018) and dried at 30 tion corresponded to a faster pH drop. Then, when Pb concentration
°C for 3 days before subjected to mass measurement. Total precipitation was over 30 mM, pH displayed a sharp drop from (9~9.25) to 8.25 in
mass was obtained by measuring mass of unused filter membrane be- 24 h and then remained almost unchanged. Fig. 2(c) shows the change
fore filtration and mass of filter membrane with precipitation settled in viable cell number in terms of CFU in solution with time. The viable
on it after drying. The filters were weighed individually before and cell number at 0 h (immediately after bacteria contacts with Pb) was be-
after filtration. All abovementioned tests had three biological replicates tween 106 and 107 CFU/mL. At Pb concentration no higher than 30 mM,
with coefficient of variance (COV) b10%. viable cell number went up to 109 CFU/mL at 24 h, indicating that bac-
teria were in the exponential growth stage. After 24 h, viable cell num-
2.4. Geochemical modelling on Pb precipitation ber remained stable, indicating that bacteria were in the stationary
growth stage. However, when Pb concentration reached 40 mM, viable
Precipitation speciation in the Pb immobilization tests was simu- cell number dropped from 107 to 104 CFU/mL at the end of the test, in-
lated using Visual MINTEQ software. Although Visual MINTEQ was un- dicating that bacteria were in the death condition. At 50 mM Pb, viable
able to simulate bacterial activities due to limitation of its database, it cells were not detected even after 12 h.
could model the evolution of precipitation speciation with increased The ability of ureolytic bacteria to induce MICP and immobilize
amount of hydrolyzed urea. More specifically in the simulation, urea hy- heavy metal contamination depended primarily on their urease activity

drolysis was modeled as an incremental titration of NH+ 4 and CO3 with a (Anbu et al., 2016). Fig. 2(d) shows the evolution of urease activity with
stoichiometric ratio of 2:1 (Gat et al., 2017), which were input as initial time. Urease activity fluctuated between 0 and 4 mM hydrolyzed urea/
conditions for various simulation cases. The initial conditions in Case 1, min in initial 4 h regardless of Pb concentration, possibly due to adaptive
Case 2 and Case 3 are summarized in Table 1. Then the evolution of var- process of bacteria after being introduced to a new medium. Then at Pb
ious precipitation phases could be obtained at different amount of hy- concentration lower than 30 mM, urease activity steadily increased to
drolyzed urea (from 0 to 200 mM). In this way, the biotic effect of 8–12 mM hydrolyzed urea/min at the end of the test. In contrast, urease
MICP on Pb precipitation was obtained. activity was mostly below 2 mM hydrolyzed urea/min at Pb
concentration ≥ 30 mM.
3. Results and analysis
3.2. Pb immobilization tests
3.1. Bacterial activity tests
Fig. 3(a) displays the change of viable cell number in solution during
Bacterial metabolic activities utilized nutrients as the energy source the Pb immobilization test. In Case 1, viable cell number at 0 h dropped
through breaking down macromolecules. This process led to an increase from 5 × 106 to 5 × 103 CFU/mL when Pb concentration increased to
in ion concentration and thereby an increase in EC value (Krishnamurti 30 mM. It then recovered slightly to 104 CFU/mL when Pb concentration
and Kate, 1951). Therefore in this test, EC value was used as an indicator reached 50 mM. However, at 24 and 48 h, viable cells could not be de-
of bacterial metabolic activity. Although the initial EC value depended tected when Pb was present in solution. In Case 2, viable cell number
on Pb(NO3)2 concentration, the further increase in EC value, more spe- steadily dropped from around 107 CFU/mL to b104 CFU/mL when Pb
cifically the EC change rate, reflected the bacterial metabolic activity. concentration reached 50 mM. Similarly as in Case 1, no viable cells
Fig. 2(a) shows the evolution of EC value with time in the bacterial activ- existed after 24 and 48 h as long as Pb contaminant was present. Finally
ity test. At Pb concentration between 0 and 30 mM, EC increased in Case 3, viable cell number maintained stable although reduction was
steadily from 9 ~ 10 mS/cm to 12 ~ 16 mS/cm in 72 h, with fastest observed at 40 and 50 mM Pb after 24 and 48 h. The significant differ-
change rate occurring at 30 mM. This indicates that bacterial metabolic ence between Case 2 and Case 3 in terms of viable cell number at 24
activity reached peak at 30 mM of Pb. However, when Pb concentration and 48 h was likely due to the amount of yeast extract presented in
was higher than 30 mM, EC increased from around 9.5 mS/cm to around the solution since yeast extract was a rich carbon source, providing
12.5 mS/cm within a very short period, but totally stopped changing amino-acids essential for bacterial growth (Bornside and Kallio, 1956).
after 8 h. This indicates that, at high Pb concentration (N30 mM), bacte- Thus, bacteria cells in Case 3 had sustained rich carbon and amino-
rial metabolic activity was intensively stimulated upon contact with Pb acid source to maintain high cell concentration.
contaminants, but diminished quickly in a few hours. Though not exper- It should be noted that there were substantial differences in terms of
imentally determined, the faster EC increase in the initial hours at viable cell number in the bacteria activity test and Pb immobilization
higher Pb concentration was likely due to higher salinity brought by test. Without direct experimental evidence, we could not definitely ex-
more Pb contaminants and stronger bacterial metabolism for plain this phenomenon. One hypothesis was that both abiotically and
biotically formed Pb precipitation in the Pb immobilization test was
Table 1 likely to encapsulate bacterial cells, cutting off their nutrient supply. En-
Initial conditions for the geochemical modelling. capsulated cells lost viability faster. Therefore, the measured viable cell
Chemical components Case 1 (mM) Case 2 and Case 3 (mM) number in the Pb immobilization test was mostly smaller than that in
2+ the bacterial activity test. Moreover, the forms of carbonate precipita-
Calcium (Ca ) 200 200
Lead (Pb2+) 0–50 0–50 tion in the two cases were also likely to be different, which might con-
Nitrate (NO-3) 0–100 0–100 tribute to different viable cell numbers in the two test (Gorospe et al.,
Chloride (Cl−) 400 0 2013).
Acetate (CH3COO−) 0 400 Fig. 3(b) presents the ammonium concentration in solution at differ-
Urea ((NH2)2CO) 0–200 0–200
ent initial Pb concentrations. As (NH4)2SO4 was originally included in
726 N.-J. Jiang et al. / Science of the Total Environment 672 (2019) 722–731

t t

t t

Fig. 2. Variables in the bacteria activity test: (a) EC, (b) pH, (c) viable cell number and (d) urease activity in solution (Number of replicates = 3; COV b 10%).

Fig. 3. Variables in the Pb immobilization test: (a) viable cell number, (b) ammonium concentration and (c) Pb removal percentage in solution. (Number of replicates = 3; COV b 10%).
N.-J. Jiang et al. / Science of the Total Environment 672 (2019) 722–731 727

the bacterial solution, there was a baseline ammonium concentration 0.40


Case 1
(around 15 mM) in the solution. Any additional ammonium concentra- 0.35 Case 2
tion above the baseline value was due to urea hydrolysis. In Case 1, the
Case 3
ammonium concentration from ureolysis at 0 h was 20 mM (the base- 0.30
line value has been subtracted) at 10 mM initial Pb. With the increase
in initial Pb concentration, the ammonium concentration from ureolysis
0.25
?
dropped steadily and reached 5 mM at 50 mM initial Pb. With time, a 0.20
general decreasing trend of ammonium concentration was observed,
which was likely due to the volatilization of ammonia under alkaline 0.15
conditions (Gat et al., 2017). Case 2 showed very similar trend as Case 0.10
1, except that the ammonium concentration from ureolysis at 0 h was
18 mM at 10 mM initial Pb and dropped steadily to 8 mM at 50 mM ini- 0.05
tial Pb. The ammonium concentration results in Case 1 and Case 2 indi-
cated that ureolysis and MICP reactions did happen initially, but soon
0.00
stopped due to insufficient cell concentration and presence of Pb. Differ- 0 10 20 30 40 50
ently in Case 3, the ammonium concentration from ureolysis at 0 h
ranged from 18 mM to 36 mM, which was not significantly influenced Pb concentration (mM)
by initial Pb concentration. Then, the ammonium concentration in-
Fig. 4. Total precipitation mass at different initial Pb concentration in the Pb
creased with time regardless of initial Pb concentration. At 48 h, the
immobilization test (Number of replicates = 3; COV b 10%).
measured ammonium concentration from ureolysis was 110 mM and
65 mM at 10 mM and 50 mM initial Pb, respectively. The results in
Case 3 indicated that substantial amount of urea was hydrolyzed and Fig. 6. In the simulation of Case 1, it can be seen that when hydrolyzed
MICP reactions did happen throughout the entire testing period. urea concentration was 0, the Pb precipitation was in the form of
Fig. 3(c) presents Pb removal percentage in solution at different Pb PbCl2, with Pb removal percentage increased from 42% to 85% when ini-
concentrations. Pb removal percentage represents Pb immobilization tial Pb concentration increased from 10 to 50 mM. The initial PbCl2 pre-
efficiency and was expressed as Eq. (6): cipitation was purely abiotic. With slight increase in hydrolyzed urea
concentration (0–17 mM), which was a biotic effect, the Pb removal
C I −C R percentage abruptly increased to N85%. The Pb precipitation also trans-
Pb removal percentage ¼  100% ð6Þ
CI formed into PbCO3 and CaCO3, with intermediate phases in transition
((PbCl)2CO3 and Pb3(CO3)2(OH)2).
where CI stands for initial Pb concentration and CR stands for remaining In the simulation of Case 2 and Case 3, as shown in Fig. 6, it can be
Pb concentration in aqueous solution. In Case 1, Pb removal percentage seen that when there was no hydrolyzed urea, the Pb precipitation
dropped from 85~95% to around 80% when initial Pb concentration in- was in the form of Pb(OH)2, with Pb removal percentage dropped
creased to 20 mM. With Pb concentration increased to 50 mM, Pb re- from 17.8% to 8.1% when initial Pb concentration increased from
moval percentage reached around 95%. Pb removal percentage at 0 h 10 mM to 50 mM. The initial Pb(OH)2 precipitation was purely abiotic.
and 24 h were close, while it was slightly lower at 48 h. It was not With moderate increase in hydrolyzed urea concentration
clear why Pb removal percentage dropped slightly with time. One hy- (12–60 mM), which was a biotic effect, the Pb removal percentage
pothesis might be that some Pb were initially subjected to bio- abruptly increased to N97%. The Pb precipitation also transformed into
absorption (accumulation of metal cations on negatively charged cell PbCO3 and CaCO3, with intermediate phase Pb3(CO3)2(OH)2 in
surface (Kotrba et al., 2011)) and were counted as immobilized specia- transition.
tion. However, with elapsed time, viable cell number declined especially In order to verify the geochemical model, measured ammonium
at high initial Pb concentrations, leading to the Pb desorption. Further concentrations from ureolysis were input into the model to calculate
studies are needed to confirm this hypothesis. In Case 2, Pb removal per- Pb2+ removal percentage in solution. The comparison with experimen-
centage declined from around 45% to around 20% at 0 h when the initial tal results is shown in Fig. 3(c). It can be seen that in all three cases, the
Pb concentration increased to 20 mM, and rebounded to around 40% simulated results were close to the ones from the experiment, indicat-
with initial Pb concentration increases to 50 mM. In Case 3, Pb removal ing the validity of the developed geochemical model. The moderate var-
percentage was close to 100% when initial Pb concentration increased to iations between simulated and experimental results were possibly due
30 mM, then decreased slightly to 95–97% when initial Pb concentration to the volatilization of ammonia and pH variation during the test.
reached 50 mM. With increasing time, Pb removal percentage slightly One thing that needs to be mentioned is that bio-sorption was also a
went up, which was different from the trend in Case 1 and Case 2. contributor to Pb removal and it was likely that larger viable cell num-
Finally, the produced precipitation mass in the three cases are ber corresponded to more Pb being absorbed biologically (Kotrba
shown in Fig. 4. It is clearly that at an identical Pb concentration, Case et al., 2011). Thus, it was rational to assume that bio-sorption played a
3 had the greatest precipitation mass while Case 2 had the least. This more important role in Case 3 than Case 1 and Case 2 due to its
conclusion was consistent with the results of Pb removal percentage overwhelmingly higher viable bacteria concentration. However, it was
discussed before. In our study, the precipitation phases were not identi- unknown to what extent the bio-sorption might impact Pb removal per-
fied experimentally due to insufficient mass of collected samples. How- centage, as it could not be individually determined in the experiment.
ever, the following geochemical model simulation results could help
understand the Pb speciation at various initial Pb concentrations in 4. Discussion
the three cases.
4.1. Effect of viable cell number on urease activity
3.3. Geochemical modelling on precipitation
Urease activity is the most critical parameter to evaluate whether
The geochemical modelling results revealed the change of Pb precip- ureolytic bacteria have sufficient capacity to trigger MICP. Urease activ-
itation phases with increased amount of hydrolyzed urea. The simula- ity is found to be dominantly controlled by viable cell number in the so-
tion result of Case 1, in which CaCl2 was used, is shown in Fig. 5. The lution during the bacterial activity test (See Fig. 7). More specifically,
results of Case 2 and Case 3, in which Ca(Ac)2 was used, are shown in substantial urease activity is only present when viable cell number is
728 N.-J. Jiang et al. / Science of the Total Environment 672 (2019) 722–731

Fig. 5. Geochemical modelling results for Case 1 at various initial Pb concentrations: (a) 10 mM, (b) 20 mM, (c) 30 mM, (d) 40 mM, (e) 50 mM.

N108 CFU/mL. This reveals how elevated Pb contamination level nega- It should be noted that previous studies also show that the activity of
tively affects urease activity of S. pasteurii by reducing the amount of vi- extracellular urease is remarkably affected by pH of surrounding envi-
able cells. This correlation is consistent with the results reported by ronment (Stocks-Fischer et al., 1999). However, in our study, the intra-
Achal et al. (2012a), with our study covers wider range of Pb concentra- cellular and extracellular urease cannot be distinguished and
tion, which is a novelty of the current study. According to Srivastava and intracellular urease activity is dependent on the viable cell number.
Kowshik (2018), heavy metal contaminants at high concentration can Thus, the urease activity is not directly linked to pH in our study.
make bacteria incapable of producing enough exopolysaccharides
(EPS) to absorb heavy metals and thereby bacteria cells are deactivated 4.2. Effect of viable cell number and calcium source on Pb immobilization
and then become dead. This leads to a reduction in viable cell number via MICP
and consequently reduced urease activity.
Mugwar and Harbottle (2016) conducted similar tests on toxicity ef- Fig. 8 shows the relationship between Pb removal percentage and vi-
fect of Pb on S. pasteurii, where 1 mM of Pb (in total concentration) was able cell number in the Pb immobilization test. It is found that Pb re-
found to suppress OD600. This threshold value was remarkably lower moval percentage has a positive relationship with viable cell number
than 30 mM in our study. On the other hand, Kang et al. (2015) showed in all three cases, though different calcium sources lead to slightly differ-
that several ureolytic bacteria strains could withstand Pb concentration ent increase pattern. This indicates that it is important to maintain suf-
up to 12.5–60 mM. It is likely that the difference in the threshold Pb con- ficient viable cells in order to reach higher Pb removal percentage,
centration is due to the variance of Pb bioavailability in different culture regardless of Pb concentration.
media. In particular, yeast extract may reduce Pb bioavailability and in- The viable cell number and calcium source change the Pb removal
crease threshold Pb concentration. percentage in three ways: abiotic precipitation, biotic precipitation
N.-J. Jiang et al. / Science of the Total Environment 672 (2019) 722–731 729

Fig. 6. Geochemical modelling results for Case 2 and 3 at various initial Pb concentrations: (a) 10 mM, (b) 20 mM, (c) 30 mM, (d) 40 mM, (e) 50 mM.

and bio-sorption, according to the experimental and geochemical simu-


lation results obtained in this study. Firstly, the calcium resource deter-
mines the initial abiotic precipitation. When CaCl2 is present, the initial
abiotic precipitation is PbCl2, which can achieve 42% to 85% Pb removal
percentage at the Pb range used in this study. When Ca(Ac)2 is present,
the precipitation is Pb(OH)2 with the Pb removal percentage ranging
from 8% to 18%. Secondly, the viable cell number determines the
ureolysis-induced and bio-sorption-induced precipitations, both are bi-
otic but cannot be separated in the current study. Therefore, in a Pb-
containing aqueous environment where substantial viable cells can sur-
vive and actively function for a long time, the biotic precipitation will be
significant, as shown in Fig. 8(a) and (b). The comparison between Case
2 and Case 3 clearly shows the difference in biotic precipitation when vi-
able cell number varies.
To the knowledge of the authors, there are few existing studies that
explicitly illustrate the relationship between Pb immobilization per-
centage and viable ureolytic bacteria cell number. Li et al. (2013) re-
ported the evolution of Pb removal rate with time when treated by
Fig. 7. The effect of viable cell number on urease activity in the bacteria activity test. different urease producing bacteria. Pb removal rate exceeded 80%
730 N.-J. Jiang et al. / Science of the Total Environment 672 (2019) 722–731

100 of numerous bacterial cells, the precipitates are likely to be nucleated


(a)
Pb removal percentage in solution (%)

based on these cells (Ferris et al., 1994, 1995). The initially formed abi-
Upper bound
otic precipitates are likely to be partially converted into biotic precipi-
tates which form the outer layer and encapsulate the initial abiotic
90 precipitates, preventing their further conversion. Accordingly, we have
proposed a hypothesized multi-layer precipitation structure, as sche-
matically illustrated in Fig. 9. In this structure, abiotic precipitates
10 mM (PbCl2 in Case 1 and Pb(OH)2 in Case 2 and Case 3) nucleate based on
80 20 mM bacterial cells and are wrapped by biotic precipitates in the sequence
Lower bound 30 mM
of (PbCl)2CO3, PbCO3, and CaCO3 in Case 1 and Pb3(CO3)2(OH)2, PbCO3
40 mM
50 mM and CaCO3 in Case 2 and Case 3. The CaCO3 on the outer perimeter can
protect Pb-related precipitations from being flushed or acid leached.
103 104 105 This hypothesized multi-layer precipitation structure is similar to the
Viable cell numbers in solution (CFU/mL) conceptual model of ureolysis driven calcite precipitation approach for
remediation of 90Sr contaminated medium proposed by Fujita et al.
(2010). Although geochemical modelling results support this hypothe-
100 (b)
Pb removal percentage in solution (%)

sis, further studies, in particular physicochemical, mineralogical and mi-


Upper bound crostructural analyses, are needed to confirm it.
80 Case 3
Case 2
60 5. Conclusions
Lower bound
40 In this study, the bacterial activity and Pb immobilization tests were
10 mM
20 mM conducted to examine the effectiveness and mechanisms of MICP for Pb
20 30 mM
40 mM
50 mM
0
101 102 103 104 105 106 107 108 109 (a)
Viable cell number in solution (CFU/mL)

C
Fig. 8. The effect of viable cell number on Pb removal percentage in the Pb immobilization

aC
test: (a) Case 1; (b) Case 2 and Case 3.

Pb

O3
(P

C
bC

O3
l) 2
after 4 h. However, the relationship between Pb removal rate and OD600

C
Pb
was not explicitly given. Achal et al. (2012) reported the different pre- Negative

O3
(C
cipitate phases of Pb when MICP was used for Pb immobilization. In par- charged
l) 2
ticular, carbonate fraction was identified as dominant precipitate.
However, biotic and abiotic precipitations were not explicitly distin-
C

guished. Fujita et al. (2010) performed a geochemical modelling to spe-


el
l

cifically reveal the evolution of different precipitate fractions of heavy


metal when subjected to ureolysis immobilization. However, the results
were only limited to 90Sr in his study.

4.3. Mechanisms of Pb immobilization via MICP


(b)
From the above discussion, it can be seen that the efficiency of Pb im-
mobilization via MICP is controlled by both abiotic and biotic mecha-
nisms. Abiotically, the low Ksp value of PbCl2 leads to its precipitation
when Cl− ions are presented in the solution in Case 1. In Case 2 and
Case 3, the initial aqueous chemical composition leads to an acid-base
C
Pb

aC

equilibrium in favor of the precipitation of Pb(OH)2. Biotically, the grad-


Pb
3

O3
(C

ual urea hydrolysis leads to the conversion of existing and formation of


C
O3
O3

new precipitates. If ureolysis can substantially happen, then the final


Negative
Pb

) 2(

precipitates are primarily PbCO3 and CaCO3, with Pb removal percent-


charged
O
(O

age nearing 100%. Meanwhile, bio-sorption, whose significance


H
H

)2

depends on viable cell concentration, can further enhance Pb


)2

immobilization.
According to the geochemical model results as demonstrated in
C
el

Figs. 5 and 6, with the increase in the degree of ureolysis, major precip-
l

itates form in the following sequence: PbCl2 → (PbCl)2CO3 → PbCO3


→ CaCO3 (Case 1) and Pb(OH)2 → Pb3(CO3)2(OH)2 → PbCO3 → CaCO3
(Case 2 and Case 3). Theoretically in an aqueous solution with high de-
gree of ureolysis, PbCl2 and (PbCl)2CO3 should be fully converted into
PbCO3 and CaCO3 in Case 1 while Pb(OH)2 and Pb3(CO3)2(OH)2 into
PbCO3 and CaCO3 in Case 2 and Case 3. However, due to the presence Fig. 9. Hypothesized multi-layer precipitation structure: (a) Case 1; (b) Case 2 and Case 3.
N.-J. Jiang et al. / Science of the Total Environment 672 (2019) 722–731 731

immobilization. Based on the results, the following conclusions can be Fujita, Y., Taylor, J.L., Wendt, L.M., Reed, D.W., Smith, R.W., 2010. Evaluating the potential
of native ureolytic microbes to remediate a 90Sr contaminated environment. Environ.
drawn: Sci. Technol. 44, 7652–7658.
Gat, D., Ronen, Z., Tsesarsky, M., 2017. Long-term sustainability of microbial-induced
(1) S. pasteurii exhibited compatible resistance to Pb toxicity when CaCO3 precipitation in aqueous media. Chemosphere 184, 524–531.
Gollapudi, U.K., Knutson, C.L., Bang, S.S., Islam, M.R., 1995. A new method for controlling
Pb concentration was no higher than 30 mM, with viable cell leaching through permeable channels. Chemosphere 30, 695–705.
number remained at 108–109 CFU/mL and urease activity at Gomez, M.G., Anderson, C.M., Dejong, J.T., Nelson, D.C., Lau, X.H., 2014. Stimulating in situ
8–12 hydrolyzed mM urea/min. soil bacteria for bio-cementation of sands. American Society of Civil Engineers
(ASCE), University of California: Davis, One Shields Ave., Davis, CA 95616, United
(2) Pb removal percentage in solution was higher at its higher initial States, pp. 1674–1682.
bacterial concentration. At identical initial bacterial concentra- Gomez, M.G., Anderson, C.M., Graddy, C.M.R., DeJong, J.T., Nelson, D.C., Ginn, T.R., 2017.
tion, removal percentage was higher when CaCl2 was used as cal- Large-scale comparison of bioaugmentation and biostimulation approaches for
biocementation of sands. J. Geotech. Geoenviron. 143.
cium source as compared to Ca(Ac)2. The highest Pb removal
Gorospe, C.M., Han, S.H., Kim, S.G., Park, J.Y., Kang, C.H., Jeong, J.H., et al., 2013. Effects of
percentage (N95%) was observed in Case 3 which had the largest different calcium salts on calcium carbonate crystal formation by Sporosarcina
viable cell number and greatest total precipitation mass. pasteurii KCTC 3558. Biotechnol. Bioprocess Eng. 18, 903–908.
(3) Ureolytic activity in the presence of Pb was largely determined Gupta P, Diwan B. Bacterial exopolysaccharide mediated heavy metal removal: a review
on biosynthesis, mechanism and remediation strategies. Biotechnology reports
by the concentration of viable cells, which were directly related (Amsterdam, Netherlands) 2017; 13.
to Pb toxicity in solution. High urease activity was only possible Jimenez-Lopez, C., Jroundi, F., Pascolini, C., Rodriguez-Navarro, C., Pinar-Larrubia, G.,
when viable cell number exceeded 108 CFU/mL. Pb immobiliza- Rodriguez-Gallego, M., et al., 2008. Consolidation of quarry calcarenite by calcium
carbonate precipitation induced by bacteria activated among the microbiota
tion efficiency was determined by abiotic precipitation, biotic inhabiting the stone. Int. Biodeterior. Biodegrad. 62, 352–363.
precipitation and bio-sorption. Abiotic precipitation was depen- Kang, C.H., So, J.S., 2016. Heavy metal and antibiotic resistance of ureolytic bacteria and
dent on calcium source and initial Pb concentration. Biotic pre- their immobilization of heavy metals. Ecol. Eng. 97, 304–312.
Kang, C.-H., Oh, S.J., Shin, Y., Han, S.-H., Nam, I.-H., So, J.-S., 2015. Bioremediation of lead by
cipitation depended on the degree of ureolysis. Bio-sorption ureolytic bacteria isolated from soil at abandoned metal mines in South Korea. Ecol.
was largely controlled by the viable cell number. Eng. 74, 402–407.
Keykha, H.A., Asadi, A., Huat, B.B.K., Kawasaki, S., 2018. Laboratory conditions for maximal
calcium carbonate precipitation induced by Sporosarcina pasteurii and Sporosarcina
aquimarina bacteria. Environ. Geotech. 1–20.
Kotrba, P., Mackova, M., Macek, T., 2011. Microbial Biosorption of Metals. Springer, New
Acknowledgements York.
This work is supported by the National Key Research and Develop- Krishnamurti, K., Kate, S.R., 1951. Changes in electrical conductivity during bacterial
ment Program of China (Grant Nos. 2018YFC1803100 and growth. Nature 168, 170.
Li, M., Cheng, X.H., Guo, H.X., 2013. Heavy metal removal by biomineralization of urease
2018YFC1802300), Environmental Protection Scientific Research Pro- producing bacteria isolated from soil. Int. Biodeterioration & Biodegradation 76,
ject of Jiangsu Province (Grant No. 2016031), National Natural Science 81–85.
Foundation of China (Grant Nos. 41472258 and 41877248), Natural Sci- Li, M.M., Fu, Q.L., Zhang, Q.Z., Achal, V., Kawasaki, S., 2015. Bio-grout based on microbially
induced sand solidification by means of asparaginase activity. Sci. Rep. 5.
ence Foundation of Jiangsu Province (Grant No. BK20170394), China Li, M., Cheng, X.H., Guo, H.X., Yang, Z., 2016. Biomineralization of carbonate by Terrabacter
Postdoctoral Science Foundation (Grant No. 2017M621595), and Indo- Tumescens for heavy metal removal and biogrouting applications. J. Environ. Eng.
U.S. Science and Technology Forum (Grant No. IUSSTF/AUG/JC/047/ 142.
Mugwar, A.J., Harbottle, M.J., 2016. Toxicity effects on metal sequestration by microbially-
2018).
induced carbonate precipitation. J. Hazard. Mater. 314, 237–248.
Mwandira, W., Nakashima, K., Kawasaki, S., 2017. Bioremediation of lead-contaminated
mine waste by Pararhodobacter sp. based on the microbially induced calcium carbon-
ate precipitation technique and its effects on strength of coarse and fine grained sand.
References
Ecol. Eng. 109, 57–64.
Nagajyoti, P.C., Lee, K.D., Sreekanth, T.V.M., 2010. Heavy metals, occurrence and toxicity
Achal, V., Pan, X.L., Fu, Q.L., Zhang, D.Y., 2012a. Biomineralization based remediation of As
for plants: a review. Environ. Chem. Lett. 8, 199–216.
(III) contaminated soil by Sporosarcina ginsengisoli. J. Hazard. Mater. 201, 178–184.
van Paassen, L.A., 2011. Bio-mediated ground improvement: from laboratory experiment
Achal, V., Pan, X.L., Zhang, D.Y., Fu, Q.L., 2012b. Bioremediation of Pb-contaminated soil
to pilot applications. ASCE, Dallas, Texas, United States 4099–4108.
based on microbially induced calcite precipitation. J. Microbiol. Biotechnol. 22,
Phillips, A.J., Cunningham, A.B., Gerlach, R., Hiebert, R., Hwang, C.C., Lomans, B.P., et al.,
244–247.
2016. Fracture sealing with microbially-induced calcium carbonate precipitation: a
Anbu, P., Kang, C.H., Shin, Y.J., So, J.S., 2016. Formations of calcium carbonate minerals by
field study. Environmental Science & Technology 50, 4111–4117.
bacteria and its multiple applications. SpringerPlus 5.
Qian, X., Fang, C., Huang, M., Achal, V., 2017. Characterization of fungal-mediated carbon-
Van Paassen, L.A., 2009. Biogrout, ground improvement by microbial induced carbonate
ate precipitation in the biomineralization of chromate and lead from an aqueous so-
precipitation. Delft University of Technology. Delft, Netherlands.
lution and soil. J. Clean. Prod. 164, 198–208.
Bornside, G.H., Kallio, R.E., 1956. Urea-hydrolyzing bacilli. I. A physiological approach to
Sanders, E.R., 2012. Aseptic laboratory techniques: plating methods. Journal of Visualized
identification. J. Bacteriol. 71, 627–634.
Experiments 63, e3064.
Castanier, S., Bernetrollande, M.C., Maurin, A., Perthuisot, J.P., 1993. Effects of microbial ac-
Srivastava, P., Kowshik, M., 2018. Mechanisms of bacterial heavy metal resistance and ho-
tivity on the hydrochemistry and sedimentology of Lake Logipi, Kenya. Hydrobiologia
meostasis: an overview. In: Donati, E.R. (Ed.), Heavy Metals in the Environment: Mi-
267, 99–112.
croorganisms and Bioremediation. CRC Press, Boca Raton, FL, pp. 15–42.
Castanier, S., Le Metayer-Levrel, G., Perthuisot, J.P., 1999. Ca-carbonates precipitation and
Stocks-Fischer, S., Galinat, J.K., Bang, S.S., 1999. Microbiological precipitation of CaCO3. Soil
limestone genesis - the microbiogeologist point of view. Sediment. Geol. 126, 9–23.
Biol. Biochem. 31, 1563–1571.
DeJong JT, Mortensen BM, Martinez BC, Nelson DC. Bio-mediated soil improvement.
Wei, B.G., Yang, L.S., 2010. A review of heavy metal contaminations in urban soils, urban
2010: 197–210.
road dusts and agricultural soils from China. Microchem. J. 94, 99–107.
Dejong, J.T., Soga, K., Kavazanjian, E., Burns, S.E., Van Paassen, L.A., Al Qabany, A., et al.,
Whiffin, V.S., van Paassen, L.A., Harkes, M.P., 2007. Microbial carbonate precipitation as a
2013. Biogeochemical processes and geotechnical applications: progress, opportuni-
soil improvement technique. Geomicrobiol J. 24, 417–423.
ties and challenges. Geotechnique 63, 287–301.
Xia, W.Y., Du, Y.J., Li, F.S., Li, C.P., Yan, X.L., Arulrajah, A., Wang, F., Song, D.J., 2019a. In-situ
Dhami, N.K., Alsubhi, W.R., Watkin, E., Mukherjee, A., 2017. Bacterial community dynam-
solidification/stabilization of heavy metals contaminated site soil using a dry jet
ics and biocement formation during stimulation and augmentation: implications for
mixing method and new hydroxyapatite based binder. J. Hazard. Mater. 369,
soil consolidaton. Front. Microbiol. 8.
353–361.
Duan, Q.N., Lee, J.C., Liu, Y.S., Chen, H., Hu, H.Y., 2016. Distribution of heavy metal pollution
Xia, W.Y., Du, Y.J., Li, F.S., Guo, G.L., Yan, X.L., Li, C.P., Arulrajah, A., Wang, F., Wang, S., 2019b.
in surface soil samples in China: a graphical review. Bull. Environ. Contam. Toxicol.
Field evaluation of a new hydroxyapatite based binder for ex-situ solidification/stabili-
97, 303–309.
zation of a heavy metal contaminated site soil around a Pb-Zn smelter. Constr. Build.
Ferris, F.G., Wiese, R.G., Fyfe, W.S., 1994. Precipitation of carbonate minerals by microor-
Mater. 210, 278–288. https://doi.org/10.1016/j.conbuildmat.2019.03.195.
ganisms - implications for silicate weathering and the global carbon-dioxide budget.
Zhao, Y., Yao, J., Yuan, Z.M., Wang, T.Q., Zhang, Y.Y., Wang, F., 2017. Bioremediation of Cd
Geomicrobiol J. 12, 1–13.
by strain GZ-22 isolated from mine soil based on biosorption and microbially induced
Ferris, F.G., Fratton, C.M., Gertis, J.P., Schultzelam, S., Lollar, B.S., 1995. Microbial precipita-
carbonate precipitation. Environ. Sci. Pollut. Res. 24, 372–380.
tion of a strontium calcite phase at a groundwater discharge zone near Rock-Creek,
Zhu, Y.Y., Ma, N., Jin, W.H., Wu, S.M., Sun, C.M., 2017. Genomic and transcriptomic insights
British-Columbia, Canada. Geomicrobiol J. 13, 57–67.
into calcium carbonate biomineralization by marine Actinobacterium Brevibacterium
Fujita, Y., Redden, G.D., Ingram, J.C., Cortez, M.M., Ferris, F.G., Smith, R.W., 2004. Strontium
linens BS258. Front. Microbiol. 8.
incorporation into calcite generated by bacterial ureolysis. Geochimica et
Cosmochinica Acta 68 (15), 3261–3270.

You might also like