Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

Engineering Properties of Biocementation

Coarse- and Fine-Grained Sand Catalyzed


By Bacterial Cells and Bacterial Enzyme
Tung Hoang, M.ASCE 1; James Alleman, M.ASCE 2; Bora Cetin, M.ASCE 3; and Sun-Gyu Choi 4
Downloaded from ascelibrary.org by Nanyang Technological University- Library on 02/04/20. Copyright ASCE. For personal use only; all rights reserved.

Abstract: Biological induced calcite precipitation is a potential method being investigated for improved soil stabilization. In terms of the
associated urea hydrolysis concept, three main strategies have been developed over the last 2 decades: (1) using live urease-producing bac-
teria, (2) using plant-extracted urease, and (3) using bacterial-extracted urease. This paper focused on evaluating the comparative benefits
of two of these methods (i.e., live bacterial cell or extracted bacterial urease methods for induced calcium precipitation) in terms of their
biocementation performance. Cell-based induced carbonate precipitation (ICP) (i.e., MICP) testing was completed on standard Ottawa
coarse-grained sand (#20/30), and bacterial-enzyme-based (i.e., BEICP) testing was conducted individually on both coarse-grained and
fine-grained (#50/70) sands. Distinctly higher unconfined compressive strength (UCS) was achieved with the BEICP method when evaluated
at similar levels of calcium precipitation. Residual permeability levels remained markedly higher after BEICP testing versus MICP. The UCS
of BEICP coarse-grained treated sand was approximately 450–1,500 kPa, whereas that of fine-grained treated sand had a notably lower range
(i.e., 250–900 kPa) when evaluated at similar levels of CaCO3 production. These results indicate that calcium carbonate content is not the sole
factor which impacts the strength of biocemented sand. Additional test-tube investigation of ICP-derived CaCO3 precipitation was used to
evaluate the chemical conversion efficiency for each method, i.e., live cells (i.e., Sporosarcina pasteurii) or bacterial-extracted urease. The
calcite precipitation ratio declined at higher substrate chemical concentrations. However, this ratio increased with higher rates of enzymatic
activity. DOI: 10.1061/(ASCE)MT.1943-5533.0003083. © 2020 American Society of Civil Engineers.

Introduction (enzyme-induced CaCO3 precipitation). These MICP and EICP


processes both use a similar mechanism, i.e., urea hydrolysis.
Ground stabilization is a fast-growing discipline in geotechnical MICP employs live bacterial cells to achieve ureolysis. EICP treat-
engineering. It includes more than 30 techniques classified within ment relies on free urease enzyme which was previously extracted
several categories, including replacement; densification through from jack beans provided by commercial sources (Nafisi et al.
dynamic energy; consolidation by dewatering; chemical grouting; 2019; Oliveira and Neves 2019).
admixture stabilization using cement, lime, or fly ash; thermal MICP-based research has demonstrated multiple soil-
stabilization; and so forth (Mitchell 1981; Terashi and Juran 2000; stabilization benefits (i.e., higher strength and stiffness, liquefaction
Chu et al. 2009). Within the last decade, a bio-based environmental resistance, and reduced permeability) of sandy soil. These outcomes
grouting method has been investigated as an additional ground have been established at varied experimental scales, ranging from
improvement technique via reduced permeability and enhanced laboratory column tests (Cheng et al. 2013; Choi et al. 2019a, b;
strength and stiffness of cohesionless soil based on both (1) pore- DeJong et al. 2006) to several higher-level experiments [100 m3
filling (bio-clogging); and (2) particle roughening plus interparticle sand (van Paassen et al. 2010a), and 1,000 m3 gravel (van der Star
bridging (biocementation) (Ivanov and Chu 2008). The majority of et al. 2011)], and even field-scale evaluation (Burbank et al. 2011;
these bio-grouting methods rely on a biochemical reaction whose Gomez et al. 2015; van Paassen 2011; Saneiyan et al. 2019).
pH-elevating chemical products then induce calcium carbonate Although the MICP method has been successful within these vari-
(CaCO3 ) precipitation. Hamed Khodadadi et al. (2017) identified ous scales of ground improvement testing, this method still has
these strategies for microbial-induced CaCO3 precipitation (MICP) several limitations because of nonuniform outcomes. Notably, prob-
as either MICP (using intact ureolytic microbial cells) or EICP lems have been experienced with both nonhomogenous distribu-
tion of bacterial cells and cementation results (Sun et al. 2018;
1
Lecturer, Univ. of Danang–Univ. of Science and Technology, Danang
Whiffin 2004; Whiffin et al. 2007), as well as constraints on the
550000, Vietnam. Email: hptung@dut.udn.vn pore-space transport of microorganism cells, in which they are re-
2
Professor, Dept. of Civil and Environmental Engineering, Univ. of strained by the presence of fine-grained soil particles (Rebata-Landa
Notre Dame, Notre Dame, IN 46556. Email: jalleman@nd.edu and Santamarina 2006). By comparison, the newer ureolysis meth-
3
Assistant Professor, Dept. of Civil and Environmental Engineering, ods using nanoscale and water-soluble urease enzyme biocatalysts
East Lansing, MI 48824 (corresponding author). Email: cetinbor@msu.edu (i.e., enzymes extracted from either plant or live cell sources) for
4
Senior Researcher, Disaster Prevention Research Division, National in situ calcium precipitation may offer beneficial processing bene-
Disaster Management Research Institute, Ulsan 44538, Republic of Korea. fits. Several researchers have published results of plant-derived
Email: choisg@kaist.ac.kr
free-enzyme biostabilization studies (e.g., Yasuhara et al. 2012;
Note. This manuscript was submitted on October 28, 2018; approved on
August 22, 2019; published online on January 22, 2020. Discussion period Hamdan et al. 2013). However, these researchers commented on
open until June 22, 2020; separate discussions must be submitted for in- the high cost of purchasing commercially purified plant-derived
dividual papers. This paper is part of the Journal of Materials in Civil urease enzyme for their lab-scale experiments (Almajed et al. 2018;
Engineering, © ASCE, ISSN 0899-1561. Hamdan and Kavazanjian 2016; Kavazanjian and Hamdan 2015;

© ASCE 04020030-1 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2020, 32(4): 04020030


Larsen et al. 2008; Nafisi et al. 2019; Nam et al. 2015; Nemati and
Voordouw 2003; Neupane et al. 2013, 2015a; Pasillas et al. 2018;
Phua and Røyne 2018; Simatupang and Okamura 2017). Currently,
only three research groups have reported on the innovative use
of self-extracted urease obtained from agricultural sources. These
studies used jack bean (Park et al. 2014) and watermelon seeds
(Dilrukshi et al. 2016, 2018; Javadi et al. 2018). Unfortunately, this
approach to extracting urease enzyme from plant sources appears
to have a number of drawbacks, including high time- and land-
demand for plant growth, plus additional high processing costs for
Downloaded from ascelibrary.org by Nanyang Technological University- Library on 02/04/20. Copyright ASCE. For personal use only; all rights reserved.

converting plant biomass into soluble extracted enzyme. To date,


very few plant-based EICP studies have compared biostabilization
efficiency relative to that of conventional MICP processing. Sim-
ilarly, only a few of these plant-based EICP studies have addressed
the effect of sand type and size (Hamdan et al. 2013; Hamdan and
Kavazanjian 2016; Kavazanjian et al. 2017; Neupane et al. 2015b;
Nafisi et al. 2019; Pasillas et al. 2018; Simatupang and Okamura
2017). Fig. 1. Grain-size distribution curve of tested sands.
Hoang et al. (2018) documented a new biogeotechnical ground
improvement method using bacterial enzyme calcium carbonate
precipitation (BEICP). The present paper further demonstrates the of filter pad material (i.e., 3M Scotch Brite Model 7447, Livonia,
performance of BEICP versus standard MICP processing method in Michigan) was placed on top of the gravel layer. Third, a moist-
terms of chemical conversion efficiency and coarse-sand stabiliza- tamping method was used to pack the specimens, in which the pre-
tion. This paper also provides further insight regarding the impact of moistened sand was tamped gently within the PVC columns (50 mm
varying sand types (coarse- and fine-grained) in relation to this new diameter and 100 mm height). To achieve a similar specific density in
BEICP strategy. The results showed similar gains in chemical con- each layer, predetermined amounts of soil were added in 10 succes-
version efficiency when using either whole-cell (MICP) or enzyme- sive layers of equal thickness (10 mm=layer) within each column.
only (BEICP) processing. However, distinct differences were found These column compaction steps were carefully conducted to achieve
when using these alternative methods with varying sand sizes. similar relative density (Dr ) levels within all column samples. The
characteristics of sample columns for coarse and fine sands including
Dr , porosity (n), and untreated permeability (k) are listed in Table 1.
Finally, another filter pad was located on the top of the sand layer to
Materials and Experimental Procedures
avoid scouring and distributing flow streams. Fig. 2 provides a sche-
matic of a filled column in operation, including its solution pump.
Sand Materials and Sand Column Preparation
Two types of silica (SiO2 )/quartz sand were used, coarse-grained
Urease-Producing Bacteria Suspension and
sand #20/30 and fine-grained sand #50/70. These sand criteria are
Chemical Solution
described in ASTM C778 (ASTM 2014). These sands contained
more than 98.7% silica. The sand materials initially were washed The urease active strain of Sporosarcina pasteurii (ATCC-11859,
with deionized water to remove any soluble chemicals, followed by American Type Culture Collection, Manassas, Virginia) was pro-
oven drying at 105°C for 24 h before being tested. Further evalu- duced as described by Hoang et al. (2018). The isolated strain was
ation of the sand properties included particle diameter at 10% finer cultivated in a presterilized culturing broth (i.e., an ammonium–
by mass (D10 ), particle diameter at 50% finer by mass (D50 ), uni- trypticase soy broth (NH4-TSB) growth media which included
formity coefficient (Cu ), coefficient of gradation (Cc ), specific trypticase soy broth (20 g=L), ammonium sulfate (10 g=L), and
gravity (Gs ), maximum and a minimum void ratio (emax and emin ). Tris buffer (0.13 mol=L); the overall solution pH was 9.0. A shak-
The sand was poorly graded sand (SP) according to the Unified Soil ing incubator (Innova Model 4000, New Brunswick Scientific, En-
Classification System (USCS) (ASTM 2010). These properties are field, Connecticut) was used to incubate this culture for 48 h with a
collectively summarized in Table 1. The grain-size distribution continuous shaking speed of 160 rpm at 30°C.
curves of the coarse and fine sands are presented in Fig. 1. This cultured biomass optical density (OD) at 600 nm was mea-
Sand columns were packed separately to include either coarse- sured using a visible light spectrophotometer (Hach Model #DR
grained or fine-grained sand soil. Prior to loading sand into each test 3900, Loveland, Colorado). When the urease-producing bacteria
column, the dry sand was mixed with distilled water to achieve 5% (UPB) solution OD was in the range 0.9–1.3, the stock microbial
moisture (Choi et al. 2019b; Hoang et al. 2018). The following se- culture was harvested and stored at 4°C. An electric conductivity
quential preparation steps were followed. First, an underlying gravel meter was used to measure the urease activity of the UPB solution
stratum with a nominal maximum size of 9.5 mm was layered into (Chu et al. 2012), which ranged from 8 to 13 mM urea=min, similar
the bottom of the test chamber’s plastic cap. Second, a precut section to that reported by Hoang et al. (2018). The concentrations of

Table 1. Index properties of sands used in the current study


Properties Column packing
Sand
type D10 (mm) D50 (mm) Cu Cc Gs emax emin USCS Method Dr (%) n (%) k (cm=s)
Coarse 0.61 0.72 1.21 1.02 2.65 0.74 0.51 SP Moist tamping 64.6  0.5 ∼37 1.01 × 10−1
Fine 0.26 0.36 1.46 0.97 2.65 0.87 0.55 SP Moist tamping 64.4  0.5 ∼40 3.2 × 10−2

© ASCE 04020030-2 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2020, 32(4): 04020030


Table 2. Test-tube experimental conditions
Number of test-tube
sample repetitious
Test substrate
(urea and CaCl2 )
concentrations
Urease
(mole)
Cells/ activitya
enzyme Test ID (mM urea/min) 0.15 0.25 0.5
Cells Cell UA-2 2 3 3 3
Cell UA-5 5 3 3 3
Downloaded from ascelibrary.org by Nanyang Technological University- Library on 02/04/20. Copyright ASCE. For personal use only; all rights reserved.

Cell UA-10 10 3 3 3
Enzyme Enzyme UA-2 2 3 3 3
Enzyme UA-5 5 3 3 3
Enzyme UA-10 10 3 3 3
a
Activity rates were derived.

three replicates were tested. The activity rates for these tests were
derived (with ∼2%–3% accuracy) by proportionate dilutions from
an initial free enzyme solution with a known (and higher) enzyme
activity. These test tubes were shaken continuously for 24 h on a
shake table at 25°C  1°C to ensure complete mixing. At the com-
pletion of this shaking period, the product solutions were filtered
Fig. 2. Soil column circulating-percolation treatment. (Reprinted from through filter paper. These tubes, as well as the filter paper samples,
Choi et al. 2016b, © ASCE.) were oven dried at 100°C  5°C for 24 h. The weight of each cal-
cium carbonate precipitation product (i.e., on the filter paper and
attached to the test tube wall) was quantified on the basis of pretest-
chemical solution used in the current study were 0.3 M, the same ing and posttesting weights. The total produced CaCO3 contents
molar concentrations as the soluble CaCl2 and urea. were calculated by adding the CaCO3 precipitated onto the test tube
surface and that remaining on the filter paper. The calcium con-
sumed efficiency was computed as the ratio of actual measured
Biomass Sonication and Enzyme Extraction
CaCO3 mass to the theoretical CaCO3 mass determined accord-
Urease enzyme lysed directly from live bacteria of Sporosarcina ing to the input chemical concentration (Neupane et al. 2013;
pasteurii was used in the present study. The procedures for extrac- Al Qabany et al. 2012).
tion and optimization were described by Hoang et al. (2018). Urease
extraction was completed using a repetitive series of cyclic run–cool Testing Program and Biotreatment Procedures
(i.e., 10 min on, followed by 10 min off) sonication steps. Six cycles
were applied over a 2-h period, with a 150-mL aliquot of the origi- Each specimen was treated with either MICP or BEICP processing
nal stock microbial culture placed directly into a sonication bath to compare the impact of these treatment methods on properties of
(Bransonic Model 220, Sterling Heights, Michigan; 120 V, 125 W, coarse- and fine-grained biocemented sand. Sand columns were
and 50=60 kHz). The cyclic processing mode was adopted to pro- treated with either 4, 8, 12, or 16 cycles because Choi et al. (2016a)
vide the lowest temperature buildup (i.e., typically at slightly below and Choi et al. (2019c) used similar levels of ICP cycling. The ex-
32°C–34°C) and highest residual enzyme activity. After completing perimental program is given in Table 3, with three replicate samples
these tests, the remaining lysed solution was subjected to a relative at each testing condition in order to verify reproducibility. Two
centrifugal force (RCF) of 5,500 for 20 min to separate out residual modes of soil stabilization were evaluated using the sand columns:
cellular debris solids from the extracted soluble urease. The sonica- (1) BEICP, and (2) MICP. A circulated-percolation down-flow pro-
tion method typically produced an approximate urease activity rate cess was applied to treat these sand columns under partially satu-
of 25.4 mM urea=min. rated conditions. The sandy-soil columns were processed using the
following sequential procedure (Fig. 2). First, the catalytic biologi-
cal solution (extracted urease for the BEICP method, and bacterial
Test-Tube Experiments
cells for the MICP method) was pumped and recycled into the top
Two different modes of test-tube experiments were used to examine of a sand column and gravity-drained from the bottom. A peristaltic
and compare calcium consumption efficiency impacted by either pump (Masterflex Model 77202-50) with silicone tubing (Master-
bacterial cells (MICP) or extracted bacterial urease (BEICP). These flex Model 96410-16) was used to recirculate this biological liquid
tests were performed within transparent 50-mL polypropylene for 3 h at a rate of approximately 5 mL=min to achieve a 60% sat-
tubes. Neupane et al. (2013), Yasuhara et al. (2012), and Carmona uration level, consistent with prior research by Cheng et al. (2013),
et al. (2018) performed similar test-tube experiments to evaluate the which allowed the bacterial cells or extracted enzyme to sorb onto
calcium carbonate precipitation ratio at different concentrations of or be trapped on the sand particle surfaces. Second, after the 3-h
commercial urease. A 15-mL equimolar solution of urea and CaCl2 introduction of the catalytic agents (i.e., either MICP culture or
was mixed with a 15-mL volume of either live bacteria cells or BEICP enzymes), the pore-volume biological liquid was drained
extracted bacterial urease. These bioreactive mixtures were evalu- off the soil column. Third, a mixed chemical solution of urea and
ated at various urease activity levels. The experimental conditions calcium chloride (0.3 M at 1∶1 ratio) was introduced and circu-
are listed in Table 2. To ensure the repeatability of these reactions, lated for 9–12 h. Fourth, the sand column was flushed with cyclic

© ASCE 04020030-3 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2020, 32(4): 04020030


Table 3. Experiment details Results and Discussion
Number of specimens
Test ICP Treatment Sand UCS Permeability CaCO3 SEM/ Whole-Cell versus Enzyme-Only Chemical Conversion
group option cycles type test test test EDS Efficiency
1 MICP 4 Coarse 3 3 3 — The chemical conversion ratios of various combinations of en-
2 8 3 3 3 1 zymatic solutions (i.e., either live cells or extracted enzyme) and
3 12 3 3 3 1 substrate reagents were evaluated using a series of test-tube experi-
4 16 3 3 3 1 ments. Fig. 3 shows the efficiencies of chemical conversion in re-
5 BEICP 4 Coarse 3 3 3 — lation to the concentration of CaCl2 for both biological sources.
6 8 3 3 3 1 In general, the chemical conversion decreased as the substrate con-
Downloaded from ascelibrary.org by Nanyang Technological University- Library on 02/04/20. Copyright ASCE. For personal use only; all rights reserved.

7 12 3 3 3 1 centration levels increased. The lowest level of precipitation effi-


8 16 3 3 3 1 ciency was at 0.5 mol=L of reagent solution.
9 BEICP 4 Fine 3 3 3 — This result agrees with those of Al Qabany et al. (2012) for
10 8 3 3 3 1 MICP and of Neupane et al. (2013) and Putra et al. (2017) for EICP.
11 12 3 3 3 1 The results by all these researchers suggested that high substrate
(i.e., urea and CaCl2 ) concentrations appeared to restrain the urease
activity of whole bacterial cells or free enzyme. A related hy-
pothesis was offered in the related literature regarding a possibly
deionized water pumping for 2 h to remove soluble byproducts and negative impact caused by thicker calcium precipitation matrixes
then the bottom cap was removed to drain off all liquid for approx- which seemingly retarded ureolytic hydrolysis (Al Qabany et al.
imately 10 h. After completing this treatment cycle (introducing 2012; Al Qabany and Soga 2013). Increased urease activity levels
urease or bacterial cells, followed by urea/calcium chloride solu- resulted in a higher efficiency of chemical conversion with both
tion addition), fresh biological solution and chemical were intro- MICP and BEICP (Fig. 3). This finding again is comparable to
duced and recirculated through the sand column in each successive that reported in the literature (Neupane et al. 2013; Putra et al.
cycle. This stepwise approach to introducing enzyme solution (or 2017). The present paper offers two different relative perspectives
bacterial cells) and urea/calcium chloride solution was repeated at on the impact of substrate concentration. At lower substrate levels
1 cycle=day for 4, 8, 12, or 16 days of total treatment phases. (i.e., 0.15 mol=L), chemical conversion efficiency increased at
higher urease activity levels. The efficiency decreased as substrate
levels increased. Although the range of the chemical efficiencies at
Testing of Properties of ICP-Cemented Sand Columns higher substrates levels was lower, they still retained a sequential
After achieving the desired cycles of treatment, each column’s correlation with enzymatic activity. Another significant aspect
bottom plastic cap and its internal filters were removed. The engi- of these test-tube studies was the observed correlation between
neering properties of the biotreated columns then were evaluated enzyme form (i.e., whole-cell versus free enzyme) and chemical
with the following laboratory experiments: permeability, unconfined conversion efficiency. The chemical conversion efficiency for
compression strength (UCS), calcium carbonate content, and micro- live cell testing was higher than that of free enzyme testing for
structure imaging. The latter testing was completed using scanning urease activity levels of 2 and 5 mM urea=min (Fig. 3). This dif-
electron microscopy (SEM) and energy-dispersive X-ray spectros- ference reversed during tests conducted at the higher activity level
copy (EDS). (i.e., 10 mM urea=min). After reaching 0.5 mol=L substrate con-
The permeability test was conducted on untreated and treated centration, though, both MICP and BEICP tests had lower effi-
sand columns. The untreated sandy soil was packed in a permeabil- ciency with calcium carbonate precipitation. Although the ureolytic
ity testing device at a similar relative density as the packed sand bacterial cells might produce more enzyme in the long term, this
columns (Table 1). The untreated sands were saturated and the per- is not expected to impact the result for chemical conversion
meability was tested following the procedure of a constant head test efficiency. Notably, other factors might restrict the activity of cells
(ASTM 2006). The permeability of biotreated specimens was
tested on the samples which were still held within the PVC test
columns. The biocemented columns were saturated by applying
15 kPa back-pressure (Cheng et al. 2013). After the initial satura-
tion step, permeability tests were run until steady permeability val-
ues (k) were reached. Tests were stopped when the k values of the
specimen did not vary by more than 20% (data not shown here for
brevity). This state is reached only when specimens are fully satu-
rated (Hoang et al. 2018).
After measuring the permeability, the columnar PVC molds
were cut to separate the biocemented samples, which were oven
dried for 48 h at a moderate temperature (i.e., ∼50°C). The UCS
test was conducted in accordance with ASTM D2324. The elastic
modulus (E50 ) of the biocementation samples was determined by
the slope of the stress–strain curve at 50% of peak stress (Oliveira
and Neves 2019; van Paassen et al. 2010a). After the samples were
broken during UCS testing, approximately 5 g of biocemented sand
was withdrawn from the middle of the sand column for calcium
Fig. 3. Chemical conversion efficiency of biological sources: bacterial
carbonate content measurement using an acid-rinsed method (Choi
cells versus bacterial enzyme.
et al. 2017b; Feng and Montoya 2015).

© ASCE 04020030-4 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2020, 32(4): 04020030


and/or free-enzymes, such as high pH (Ciurli et al. 1996; Stocks- filtering and trapping of these intact cells within the sand pore
Fischer et al. 1999) or surficial coating by calcite precipitation, matrix during down-flow percolation, in which both then induce
which then might retard the nucleation rate of calcite crystal for- CaCO3 nucleation (Ginn et al. 2002). Furthermore, it is plausible
mation (Mitchell and Ferris 2005, 2006). Therefore, actual mea- that increased numbers of treatment cycles might increase ionic par-
surements of chemical conversion were lower than the expected ticle charging, which then could boost bacterial adhesion (Faibish
theoretical chemical conversions. et al. 1998; Foppen and Schijven 2006) on sand particles in relation
The highest efficiency of chemical conversion was observed to an elevated zeta potential (Dick et al. 2006). In contrast, the lower
with 10 mM urea=min at a substrate level of 0.15 mol=L. However, levels of CaCO3 precipitation in the BEICP-treated samples may
the 10%–20% increase in the precipitation ratio corresponded to have been caused by soluble urease enzyme flushing and loss during
a 50% activity increase (i.e., from 5 to 10 mM urea=min). The cyclic drainage, as opposed to surficial binding or trapping reten-
approximate urease activity of 5 mM urea=min subsequently was tion. Back-calculated estimates of residual urease activity remaining
Downloaded from ascelibrary.org by Nanyang Technological University- Library on 02/04/20. Copyright ASCE. For personal use only; all rights reserved.

chosen for future tests given the pragmatic premise that a lower in the column, versus that which was removed via flushing, were
urease dosing level is more economical when the original bio- approximately 40% of the original rate during BEICP processing for
logical solution is to be diluted. In addition, Cheng et al. (2017) one to four treatment cycles (data not shown). However, this per-
mentioned that a low-urease-activity solution is more effective in centage of urease activity retention reached 60%–70% as the num-
improving mechanical strength. A low concentration of substrate ber of treatment cycles increased (data not shown), likely due to
reagent (0.15 mol=L) was used for the chemical solution in the col- decreased permeability (and additional enzyme trapping) which de-
umn test because it resulted in higher strength and more-uniform veloped at higher treatment cycles. Although the CaCO3 content of
samples for a given amount of calcium carbonate precipitation BEICP-treated sand did reach ∼8% after 16 cycles, this level still
(Choi et al. 2019a; Al Qabany and Soga 2013). was far lower than that of MICP-treated sands.
In general, therefore, when considered on the basis of compa-
Whole-Cell versus Enzyme-Only Impact on rable levels of calcium carbonate precipitation, the strength range
Biostabilized Coarse-Sand Properties of BEICP-treated sand tended to be higher than that of MICP-
treated sand. For example, at 2%–4% and 5%–8% levels of CaCO3 ,
The coarse-grained sand columns were stabilized via intact bacteria the UCS of BEICP-treated sand was approximately double that
and extracted enzyme ICP methods which received 4, 8, 12, or 16 achieved with MICP processing. Although Zhao et al. (2014) also
repetitive treatment cycles to achieve different levels of CaCO3 reported that EICP treatment had a lower range of CaCO3 content,
content. The properties of the biocemented samples were tested to their tests gave lower UCS results than for MICP processing. How-
comparatively evaluate the resulting variations with these alterna- ever, they used a different treatment method, i.e., without adding
tive methods in terms of UCS, elastic modulus (E50 ), permeability, fresh biological solution, so ongoing degradation of the originally
and microstructure of biocemented coarse-grained sand. added enzyme might have decreased their EICP product outcome
below that of MICP.
UCS and Elastic Modulus Although the amount of CaCO3 precipitation undoubtedly is an
Fig. 4 identifies the peak stress levels of MICP-treated samples; important factor impacting biocemented sands, the distribution pat-
these results varied considerably, from 200 to 2,300 kPa. However, terns of the precipitated crystal clusters also might have played a
the UCS estimates for BEICP-treated sands ranged from 400 to
vital role in product strength (Cheng et al. 2017; Cui et al. 2017;
1,500 kPa and, in one sample, up to approximately 2,400 kPa.
Hoang et al. 2018; Lin et al. 2016; Terzis and Laloui 2018). These
Although the UCS range of treated sand was similar for both bio-
distribution patterns involve three apparent factors: (1) location of
treatment methods, the levels of CaCO3 precipitation were substan-
the calcium crystals relative to sand interparticle contact; (2) thick-
tially different in the MICP- and BEICP-treated samples. These
ness of the crystal clusters; and (3) nonbeneficial precipitation of
precipitation values ranged from 2.5% to 16% CaCO3 for MICP-
CaCO3 within sand surfaces. Microstructure analysis provided fur-
treated samples, and from 1.5% to 8% for BEICP-treated specimens.
ther insight into the influence of these latter three factors, and these
The result demonstrated that MICP processing consistently pro-
details are presented subsequently.
duced a higher amount of CaCO3 for the same number of treatment
Fig. 5 compares the present and previously reported UCS results
cycles compared with BEICP treatment. The increased biocemen-
for CaCO3 precipitation, including the present BEICP results and
tation may be caused by whole-cell attachment to sand surfaces, and
the present and prior MICP results. The UCS strength levels re-
corded during this study are in exceptionally good agreement with
previously published MICP findings (Choi et al. 2016b, 2017a;
van Paassen et al. 2010a; Al Qabany and Soga 2013) from studies
which employed a similar substrate concentration (i.e., 0.3 M for
both urea and CaCl2 ). Other UCS data of MICP-treated coarse sand
with 1 M chemical concentration and partial saturation treatment
process by Cheng et al. (2013) are included in Fig. 5 for compari-
son. However, the UCS trend line obtained from BEICP-treated
sand in the present work was recorded at a noticeably higher range
than that of MICP-treated data in terms of comparable calcium dep-
osition levels.
Typical stress–strain curves of biocemented sand obtained from
unconfined compression testing at four and eight cycles of treat-
ment are presented in Fig. 6. Both MICP- and BEICP-treated sand
exhibited a brittle nature comparable to that observed in prior MICP
(Bernardi et al. 2014; Choi et al. 2016a; Mujah et al. 2019) and
Fig. 4. UCS results for MICP- and BEICP-treated samples of coarse
EICP (Kavazanjian and Hamdan 2015; Neupane et al. 2015a; Park
sand.
et al. 2014) testing. The peak stress of MICP-treated samples after

© ASCE 04020030-5 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2020, 32(4): 04020030


Downloaded from ascelibrary.org by Nanyang Technological University- Library on 02/04/20. Copyright ASCE. For personal use only; all rights reserved.

Fig. 7. Elastic modulus results for MICP- and BEICP-treated samples


of coarse sand.

sand specimens (Fig. 7). The increasing trend of E50 correlated to


CaCO3 content provided by MICP-treated samples was comparable
Fig. 5. UCS of present BEICP and MICP results versus previous MICP
data.
to that observed for BEICP treatment. Compared with previous
E50 test results from MICP-treated sand, van Paassen et al. (2010a)
found that E50 results with ICP-treated sand ranged from 100 to
8,500 MPa, significantly higher than the elastic modulus results
four-cycle treatment clearly was lower than that of similarly treated in the present study. However, BEICP processing produced an
BEICP-treated sand (Fig. 6). A lack of uniformity within MICP- even higher elastic modulus than the EICP results reported by
treated sands resulted in lower UCS strengths. During these sample Yasuhara et al. (2012) (50–160 MPa) and Neupane et al. (2015a)
tests, failure did not occur within the whole core. Instead, sample (16–18 MPa). van Paassen et al. (2010a) reported the use of a high
failure occurred within the less solidified and weaker bottom sec- substrate concentration (1 M CaCl2 ∶urea) and a high level of treat-
tion. Similar outcomes were reported by Cheng et al. (2013) and ment (16 days), whereas fewer treatment cycles were injected by
van Paassen et al. (2010b). Lower strengths of the MICP-treated Yasuhara et al. (2012) (4 and 8) and Neupane et al. (2015a) (2).
samples similarly could have been caused by nonuniform bioce- These data, both past and present, strongly suggest that substrate
mentation conditions after low cycles of treatment (i.e., with lower strength and number of treatment cycles are the main factors for the
calcium cementation at lower depths within treated columns). finished strength of biocemented sand.
The modulus was computed as a secant elastic modulus based
on the strain required to mobilize 50% of the peak stress, E50 . For Permeability
example, the maximum stress of BEICP-treated sand with four The permeability results for biocemented sand, using both MICP
treatment cycles was 490 kPa at 0.29% strain (Fig. 6). At this strain and BEICP treatment, are graphed in Fig. 8. The findings are again
level the E50 of the specimen corresponds to 119.8 MPa. The elastic depicted in relation to calcium precipitation levels. In general,
modulus of the MICP-treated samples ranged from 20 to 250 MPa, the ICP-treated sand samples exhibited a reduction in permeabil-
whereas it varied between 50 and 200 MPa for the BEICP-treated ity in relation to increases in CaCO3 content. The permeability
levels for BEICP-treated sands were slightly lower in the range of
1.5%–4% CaCO3 precipitation. For BEICP processing, the highest

Fig. 6. Typical stress–strain relationship for MICP- and BEICP-treated Fig. 8. Permeability results for MICP- and BEICP-treated samples of
samples of coarse sand. coarse sand.

© ASCE 04020030-6 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2020, 32(4): 04020030


reduction of permeability occurred at an 8% CaCO3 content. The precipitation, induced by either whole cells with MICP or extracted
impact of MICP processing was more pronounced, with a 3- to enzymes with BEICP, plays a crucial role in reducing the per-
4-log decrease within a 13%–16% CaCO3 precipitation range. meability of biocemented sand.
Although MICP exhibited far lower permeability reductions than
did BEICP, both treatment options had a similar pattern of reduction Microstructural Analyses
relative to CaCO3 precipitation. This similarity may be attributable The microscale morphology of precipitated calcite crystals within
to multiple factors. First, the sand columns were packed with the the coarse-grained biocemented sand matrix was analyzed using
same type of sand (coarse-grained sand, #20/30) and with the same scanning electron microscopy. Photographs obtained with the SEM
porosity (Table 2), which in turn would have been expected to evaluation are provided in Fig. 9. Figs. 9(a–c) show that calcium
produce similar initial permeabilities. Second, sample permeabil- carbonate precipitated via bacterial cells (i.e., MICP) agglomerated
and formed large crystal clusters with a substantial thickness of
Downloaded from ascelibrary.org by Nanyang Technological University- Library on 02/04/20. Copyright ASCE. For personal use only; all rights reserved.

ity would have been reduced with progressive calcium-rich crystal


formation within void spaces between sand grains. Third, escalat- 50–100 μm. One important feature in this regard was that of the
ing CaCO3 precipitation within sand pore space would have led to large CaCO3 clusters precipitated at the contact points between
the lower permeability levels within the specimens. A consequent sand particles, which accumulated within and progressively filled
observation was that progressively increased calcium carbonate interstitial gaps between the adjacent sand particles and through

Fig. 9. SEM images of bio-treated samples for coarse sand: (a–c) MICP at 12-cycle levels; and (d–f) BEICP at 12-cycle levels.

© ASCE 04020030-7 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2020, 32(4): 04020030


which agglomeration then clogged pore spaces within the sand permeability produced by MICP and BEICP. This correlation in
matrix. The crystal cluster patterns observed in this study agree shown in the trends for permeability reduction in Fig. 8. The visual
with observations reported by previous researchers (Cheng et al. evidence in Fig. 10 confirmed that far higher levels of deposited
2013, 2017; Cui et al. 2017; Mujah et al. 2019). In contrast, the calcium were attached to sand grain surfaces and pore-space vol-
small crystal clusters produced via bacterial enzyme (i.e., BEICP) umes during MICP processing [Fig. 10(b)] versus BEICP process-
processing accumulated as noticeably smaller calcium-bearing crys- ing [Fig. 10(e)]. Similar observations were made by previous
tals [Figs. 9(d–f)]. The average size of these smaller BEICP-based researchers (i.e., Cheng et al. 2013; Al Qabany and Soga 2013).
cluster ranged from 5 to 20 μm. These smaller BEICP-derived crys- One further noteworthy observation is the significant difference
tal clusters were observed primarily at the sand-grain contact points between MICP [Fig. 10(c)] and BEICP [Fig. 10(e)] processing
and only intermittently on the sand particle surfaces. Because of the in terms of their relative chloride deposition levels. Although both
smaller size of these crystals, less pore volume was filled within the modes of treatment involved tap-water flushing within 2 h after
Downloaded from ascelibrary.org by Nanyang Technological University- Library on 02/04/20. Copyright ASCE. For personal use only; all rights reserved.

BEICP-treated sands. every treatment cycle to remove residual chemicals, Fig. 10(e)
These observed differences in the size of ICP-derived calcium clearly shows a substantial residual presence of chloride. This
formation between the MICP and BEICP processes were explained chloride residual was attributed to the far lower permeability of
by Hoang et al. (2018). Most calcium crystal sizes ranged from 5 to MICP-treated samples after multiple treatment cycles. The reduced
10 μm for MICP-treated sand [Fig. 9(c)], whereas BEICP crystals permeability means that chloride likely was not flushed from the
were noticeably smaller, 1–4 μm [Fig. 9(f)]. The premise behind sample during the postprocessing tap-water rinse step, and then was
this change in crystal precipitation size is that of the nucleation baked into a solid, chloride-rich deposit inside the MICP cores
sources, derived from either whole-cell or enzyme-only mecha- during the final oven drying. This covert chloride crystallization
nisms. Mitchell and Ferris (2006) employed urea hydrolysis using behavior could well have contributed to a falsely higher measure-
whole-cell S. pastuerii bacteria for MICP processing, and observed ment of MICP mechanical strength levels compared with what
calcium crystal diameters of 4.2 and 7.4 μm in 1 and 7 days, re- would have been expected had long-term soaking and flushing
spectively. In contrast, nanosized crystal formation (∼20 nm) was of the chloride been achieved. This unexpected MICP finding sug-
reported by Sondi and Salopek-Sondi (2005), who used enzyme- gests yet another benefit of BEICP processing, in where no such
only ureolytic treatment for ICP precipitation. These findings sug- chloride deposition occurred because tap-water flushing was far
gest that the mode of urea hydrolysis is a key factor affecting the more effective due to the higher residual permeability.
size of calcite crystals in the ICP process. Perspectives regarding
the distribution patterns of calcium crystal clusters have been re- Effect of Sand Grain Size on BEICP-Treated Samples
ported for MICP processing under various conditions. Al Qabany
The following details and discussion address the influence of sand
and Soga (2013) and Choi et al. (2019c) reported the impacts of
particle size on BEICP-treated columns. The type and size of sand
substrate concentrations, retention times, and chemical addition
grains subjected to ICP processing are key factors in terms of bio-
flow rates. Cheng et al. (2013, 2017) investigated the effect of vary-
cementation success. Previous researchers addressed the effects of
ing degrees of saturation, levels of urease activity, processing tem-
sand grain size ranging from fine to coarse using MICP processing
peratures, and other environmental factors. To further evaluate the
(Bernardi et al. 2014; Cheng et al. 2013, 2017; Kaur et al. 2016; Lin
distribution patterns of crystal clusters of MICP versus BEICP
et al. 2016; Martinez et al. 2013; Mortensen et al. 2011; Mwandira
processing, therefore, EDS overlay imaging was conducted for both
et al. 2017; Zhao et al. 2014; Rebata-Landa and Santamarina
sample specimens following 16 cycles of treatment (Fig. 10).
2006; Terzis and Laloui 2018, 2019; Weil et al. 2018) and EICP
Although the sand-particle contact points were bridged by
processing (Hamdan et al. 2013; Kavazanjian and Hamdan 2015;
BEICP-deposited crystals, the pore space of the sand matrix re-
Nafisi et al. 2019; Neupane et al. 2015b). The present paper adds
mained fairly open [Fig. 10(a)]. However, MICP processing pro-
further insight to this prior understanding, in terms of BEICP’s per-
duced a noticeably higher level of calcium deposition [Fig. 10(b)].
formance with coarse and fine sand. As with the previous tests of
Fig. 10 shows calcite clusters deposited at three locations: (1) sand
coarse-grained sand, these BEICP fine-sand tests were conducted
particle contact points, (2) sand grain surfaces, and (3) sand pore-
with 4, 8, and 12 cycles. Figs. 11 and 12 show the resultant UCS
space volume. These findings regarding differences in calcium
and elastic modulus. Fig. 13 further compares UCS results for
precipitation via MICP and BEICP processing appear to directly
BEICP with other published EICP results. In addition, Fig. 14
impact the finished mechanical properties in terms of strength
shows permeability results and Figs. 15 and 16 provide a set of
and stiffness of biocemented sand. One such apparent correlation
SEM images.
is that the strength of biocemented sands was more influenced by
the location of calcium crystal deposition than by the total mass UCS and Elastic Modulus
of CaCO3 buildup. This premise agrees with previous studies by The results obtained with BEICP-based treatment revealed a
Cheng et al. (2013), Cui et al. (2017), Al Qabany and Soga (2013). more pronounced strength (Fig. 11) and elastic modulus (Fig. 12)
For example, to reach 1,500 kPa of UCS, BEICP-treated sam- stabilization when applied to cohesionless materials. The UCS
ples needed approximately 6%–8% CaCO3 precipitation, whereas of BEICP-treated coarse-grained sand was between 450 and
MICP-treated sands required about 10%–11% CaCO3 content to 1,500 kPa for calcium carbonate precipitation from 2% to 6%.
achieve similar mechanical strength (Fig. 4). Therefore, in terms of BEICP processing of fine sand provided ranges of UCS from
strength achieved per mass of deposited calcium, the efficacy of 200 to 900 kPa at similar CaCO3 contents. Similar patterns of
BEICP processing appear to be higher than that of MICP treatment. increased biocemented stiffness were observed in relation to in-
Furthermore, a more specific rationale for the high efficiency of creased calcium carbonate deposition for both sand types. The
increased UCS results with BEICP processing is that calcium de- Young’s modulus of coarse biocemented sand varied between 75
posited at the contact points of sand particles governs the strength and 125 MPa, whereas that of fine biocemented sand varied from
of sand. Similar perspectives were presented by Cheng et al. (2017), 25 to 75 MPa when correlated with ranges of CaCO3 varying from
Martinez and DeJong (2009). 2% to 6% for both materials. The trend toward lower UCS and
Another related aspect of this behavior is that the amount quan- stiffness for BEICP-treated fine sand compared with coarse sand
tity of calcium deposition had a major impact on the reduction of at a similar level of CaCO3 content was consistent with results

© ASCE 04020030-8 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2020, 32(4): 04020030


Downloaded from ascelibrary.org by Nanyang Technological University- Library on 02/04/20. Copyright ASCE. For personal use only; all rights reserved.

(a) (d)

(b) (e)

(c) (f)

Fig. 10. SEM and EDS images of bio-treated samples for coarse sand: (a–c) MICP at 16-cycle levels; and (d–f) BEICP at 16-cycle levels.

Fig. 11. UCS results of BEICP-treated samples for coarse- and fine- Fig. 12. Elastic modulus results of BEICP-treated samples for coarse-
grained sands. and fine-grained sands.

© ASCE 04020030-9 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2020, 32(4): 04020030


moisture retention is higher in this hinge area under partially satu-
rated–column conditions. Although they did not offer any further
hypotheses about the extent of hinge development and correspond-
ing calcium deposition, their paper’s findings suggest that higher
calcium deposition, higher hinge volume, and higher moisture re-
tention were all more prominent when dealing with fine-grain sand.
They also indicated that for a partially saturated treatment, the cal-
cium carbonate crystal clusters precipitated mainly at the contact
points of grains where solution menisci layers formed. Therefore,
the CaCO3 contents were higher in the fine BEICP-treated sand
Downloaded from ascelibrary.org by Nanyang Technological University- Library on 02/04/20. Copyright ASCE. For personal use only; all rights reserved.

which was treated by the partial saturation method (percolation–


circulation technique).
Terzis and Laloui (2018) investigated crucial microscopic char-
Fig. 13. UCS of present BEICP results versus previous EICP data. acteristics of biocemented sand in relation to apparent biostabi-
UCS was compared based on the total precipitated CaCO3 content. lization strength, covering such factors as the particle sizes of the
crystalline bond lattice, bond-grain contacts, and particle orienta-
tions. Their study indicated that larger-sized biocemented sands
had higher strength and stiffness than smaller-sized MICP-treated
sand at a similar CaCO3 content. According to Terzis and Laloui
(2018), MICP-treated large sand had larger mean diameters of
crystal clusters precipitated at sand grain contact points than that
achieved with biocemented smaller sand. They concluded that this
biocemented coarse sand grain behavior might lead to a higher
resistance to particle shearing and increased particle interlocking.
In addition, the spatial orientation and population of calcium crys-
tal clusters in large biocemented sands were found to be more
homogenously distributed than calcium bonds precipitated with
smaller-grained sands. As a more homogenous distribution of cal-
cium crystals developed within the spatial orientation space, a
higher overall stress resistance developed due to intergranular con-
tact and cementation. Although the UCS and Young’s modulus of
coarse-grained BEICP-treated sand were higher than those of fine-
grained biocemented sand at a similar level of CaCO3 , further stud-
ies of stress–strain behavior at various confining stress, friction
Fig. 14. Permeability results of BEICP-treated samples for coarse- and angles, and cohesion should be investigated for both sand materials.
fine-grained sands. Fig. 13 compares this study’s BEICP UCS results with re-
sults of prior investigators using EICP biostabilization methods
(Kavazanjian and Hamdan 2015; Neupane et al. 2015b; Oliveira
et al. 2016; Park et al. 2014; Yasuhara et al. 2012). In this case,
previously reported by Gomez et al. (2013), Lin et al. (2016), Zhao the present study’s BEICP-derived UCS results were considerably
et al. (2014), and Terzis and Laloui (2018) for MICP processing higher than the previously published EICP findings at comparable
and by Hamdan et al. (2013) and Kavazanjian and Hamdan (2015) calcium deposition levels.
for EICP processing. In contrast, Cheng et al. (2013) mentioned
that at a similar CaCO3 content, fine MICP-treated sand achieved Permeability
higher values of cohesion and friction angle than did MICP-treated The permeability results for coarse-grained and fine-grained
coarse sand. BEICP-based sand treatment are given in Fig. 14, including values
The overall bulk mass of calcium carbonate precipitation could for both untreated materials and biostabilized samples. As expected,
not be considered the sole factor governing the level of stabilization a permeability reduction realized during BEICP sand processing
for biocemented sand [Lin et al. (2016], Terzis and Laloui (2018)). correlated with an increase in CaCO3 content for both fine and
Overall, the UCS of BEICP-treated coarse sand tended to be higher coarse sand. However, the trend of reduced permeability observed
than that of BEICP-treated fine sand at comparable levels of cal- with fine-grained biocemented sands steadily declined with higher
cium carbonate content. For example, at 2%–3% of CaCO3 precipi- calcium deposition levels, whereas the trend for coarse-grained sand
tation, the UCS values of coarse-grained bio-treated sand ranged had much less reduction in permeability as calcium carbonate pre-
from 450 to 700 kPa, whereas the UCS of fine-grained bio-treated cipitation increased. This trend of lower permeability in fine-grained
sand ranged from 380 to 540 kPa. In addition, at 5%–6% levels bio-treated sand compared with coarse-grained sand was consistent
of CaCO3 , the strength of biocemented coarse sand was approxi- with the data of Cheng et al. (2013) from MICP treatment. Fine-
mately twice the UCS of BEICP-treated fine sand. The fact that grained BEICP-treated sand typically had lower permeability values
finer sand had higher CaCO3 content and lower strength matches than the coarse-grained BEICP-treated sands at similar levels of
observations made by previous studies on MICP and EICP process- CaCO3 content. In addition, both materials were packed at a similar
ing (Kavazanjian and Hamdan 2015; Lin et al. 2016; Terzis and relative density in the untreated state. This indicates that the per-
Laloui 2018). meability reduction of BEICP-treated coarse and fine sand might
Cheng et al. (2013) characterized the nature of this same behav- have been controlled by factors other than average or bulk CaCO3
ior on the basis of a so-called hinge mode of calcium deposition at content (e.g., the distribution patterns of calcium carbonate pre-
the point of grain contact. Their corresponding assumption was that cipitation). The SEM analysis in the next section provides a visual

© ASCE 04020030-10 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2020, 32(4): 04020030


Downloaded from ascelibrary.org by Nanyang Technological University- Library on 02/04/20. Copyright ASCE. For personal use only; all rights reserved.

Fig. 15. SEM images of BEICP-treated samples: (a–c) coarse-grained sand at 8-cycle levels; and (d–f) fine-grained sand at 8-cycle levels.

perspective of calcite crystal distribution patterns observed with across the grain surfaces and in a retained menisci layer formed due
both coarse- and fine-grained BEICP-treated sand. to surface tension at the point of sand grain contact. The size of a
single calcium crystal initially formed by BEICP processing tended
Microstructural Analyses to be quite small, ranging from ∼1 to 4 μm, for both bio-treated
Figs. 15 and 16 provide a set of SEM images for BEICP-treated coarse and fine sands [Figs. 15(c and f)]. This is because the pre-
coarse-grained and fine-grained sands after 8 and 12 treatment cipitation particle was built up from nanosized crystals generated
cycles, respectively. These images depict calcium cluster deposi- from BEICP’s nanoscale free enzyme catalyst.
tion occurring at the contact points and surface of sand grains, but However, the amount and distribution pattern of calcium crystal
pore-space deposition with these BEICP samples is noticeably less precipitation differed widely between the coarse and fine materials.
than that with MCIP samples [Figs. 9(a) and 10(a)]. A similar pat- The CaCO3 levels with #50/70 sand samples [Figs. 15(d–f)] likely
tern of reduced pore-space volume deposition was reported by were higher than those in #20/30 sand [Figs. 15(a–c)]. As was dis-
Cheng et al. (2013) for MICP biostabilization applied in a partially cussed previously, the average mass of CaCO3 was not the sole
saturated state. During this sort of unsaturated processing con- factor governing the strength of biocemented sand. The calcium
ditions, and with pore-space volume primarily filled with air, the carbonate distribution pattern was another main factor influencing
down-flowing substrate solution is retained as a moist film spread the processed sample strength. The crystal clusters concentrated

© ASCE 04020030-11 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2020, 32(4): 04020030


Downloaded from ascelibrary.org by Nanyang Technological University- Library on 02/04/20. Copyright ASCE. For personal use only; all rights reserved.

Fig. 16. SEM images of BEICP-treated samples: (a and b) coarse-grained sand at 12-cycle levels; and (c and d) fine-grained sand at 12-cycle levels.

mainly at the contact points of sand grains in coarse-grained BEICP- were evaluated using BEICP. The following key conclusions were
treated sands, whereas these clusters predominantly coated the grain identified from this research effort:
surfaces of fine-grained sands. Cheng et al. (2013), Cui et al. (2017), • Chemical conversation efficiency for both MICP and BEICP
Hoang et al. (2018), Lin et al. (2016), Terzis and Laloui (2018) methods decreased as the concentration of applied substrate
all mentioned that surface-coating calcium crystals tend to provide chemicals increased. However, in both cases the chemical con-
noneffective grain bonding (i.e., compared with effective interpar- version efficiency increased as the urease activity level increased.
ticle bridging bonds, whose effectiveness likely governs stabiliza- MICP processing (using whole-cell enzymes) produced higher
tion strength). Therefore, the UCS and Young’s modulus of the fine levels of chemical conversion efficiency when evaluated at ur-
BEICP-improved sands were lower than those of coarse-grained ease activity levels between 2 and 5 mM urea=min. However, the
sands (Figs. 11 and 12). chemical conversion efficiency for BEICP (using bacterial-
The distribution pattern of calcium clusters likely impacted the extracted free enzyme), was higher than that of MICP processing
reduction of permeability in BEICP-treated sands. Fig. 15 shows the at a 10 mM urea=min rate.
clusters deposited almost exclusively at the contact points of sand • BEICP processing demonstrated higher UCS levels than MICP
grains and sand surfaces after 8 cycles, whereas there was continued with coarse sands and published EICP results at comparable cal-
deposition both at the contact points and in the internal void space cium deposition levels. BEICP processing also generated smaller
after 12 cycles (Fig. 16). However, after 12 cycles of treatment, the calcium crystal sizes at comparable levels of treatment cycles.
amount of calcium crystal filling the void space during fine-sand For example, BEICP completed at an eight-cycle treatment
BEICP treatment [Figs. 16(c and d)] likely was higher than that tak- mode produced crystals of 1–4 μm. MICP produced a noticeably
ing place within coarse-grained sands [Figs. 16(a and b)]. In this higher level of calcium deposition within sand pore-space vol-
case, therefore, more pore-volume filling, and a resultant further re- ume than at the sand surface and meniscus areas, whereas BEICP
duction in permeability, is expected. tended to realize a higher proportion of calcium deposition at
sand contact points and adjacent meniscus zones.
• MICP produced a higher reduction in permeability because of a
Conclusions and Recommendations tendency toward higher pore-space calcium deposition. Conver-
sely, BEICP processing tended to have a lower impact on redu-
This paper provided a comparative investigation of whole-cell cing permeability. This particular behavior may be viewed as a
(i.e., MICP) and extracted bacterial enzyme (i.e., BEICP) methods distinct benefit, because multiple treatment cycle applications
focusing on two key factors: (1) chemical conversion efficiency, would not be as prone to column clogging.
and (2) engineering properties. Both MICP and BEICP assess- • The observed correlation between sand grain size, levels
ments were conducted using coarse sands, whereas only fine sands of CaCO3 precipitation, and distribution patterns of calcium

© ASCE 04020030-12 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2020, 32(4): 04020030


deposition clustering were prominent factors in relation to of lignocellulosic biomass.” ACS Sustainable Chem. Eng. 5 (6): 5183–
strength and permeability outcomes. BEICP processing of fine- 5190. https://doi.org/10.1021/acssuschemeng.7b00521.
grained sands produced lower UCS levels compared with coar- Choi, S. G., J. Chu, and T. H. Kwon. 2019c. “Effect of chemical concen-
segrained sands at similar levels of CaCO3 precipitation. This trations on strength and crystal size of biocemented sand.” Geomech.
pattern for BEICP was consistent with prior EICP publications Eng. 17 (5): 465–473. https://doi.org/10.12989/gae.2019.17.5.465.
and prior MICP publications. Choi, S.-G., S.-S. Park, S. Wu, and J. Chu. 2017b. “Methods for calcium
carbonate content measurement of biocemented soils.” J. Mater. Civ.
Two recommendations are given for future research regarding
Eng. 29 (11): 06017015. https://doi.org/10.1061/(ASCE)MT.1943
the mechanisms and performance of BEICP processing. First, test-
-5533.0002064.
ing should be conducted to specifically quantify protein mass levels
Choi, S.-G., K. Wang, and J. Chu. 2016b. “Properties of biocemented, fiber
when using extracted free-enzyme removed from viable ureolytic reinforced sand.” Constr. Build. Mater. 120 (Sep): 623–629. https://doi
cells. This approach would provide an enhanced means of com-
Downloaded from ascelibrary.org by Nanyang Technological University- Library on 02/04/20. Copyright ASCE. For personal use only; all rights reserved.

.org/10.1016/j.conbuildmat.2016.05.124.
pleting a specific characterization of enzymatic reactions relative Chu, J., V. Stabnikov, and V. Ivanov. 2012. “Microbially induced calcium
to chemical conversion efficiency. Second, further investigation is carbonate precipitation on surface or in the bulk of soil.” Geomicrobiol.
warranted to qualify the use of BEICP processing within complex, J. 29 (6): 544–549. https://doi.org/10.1080/01490451.2011.592929.
natural soil systems. Chu, J., S. Varaksin, U. Klotz, and P. Menge. 2009. “Construction proc-
esses.” In Proc., 17th Int. Conf. on Soil Mechanics and Geotechnical
Engineering: The Academia and Practice of Geotechnical Engineering.
References Amsterdam, Netherlands: IOS Press.
Ciurli, S., C. Marzadori, S. Benini, S. Deiana, and C. Gessa. 1996. “Urease
Almajed, A., H. Khodadadi, and E. Kavazanjian Jr. 2018. “Sisal fiber from the soil bacterium Bacillus pasteurii: Immobilization on
reinforcement of EICP-treated soil.” In Proc., IFCEE 2018, 29–36. Ca-polygalacturonate.” Soil Biol. Biochem. 28 (6): 811–817. https://doi
Reston, VA: ASCE. .org/10.1016/0038-0717(96)00020-X.
Al Qabany, A., and K. Soga. 2013. “Effect of chemical treatment used in Cui, M. J., J. J. Zheng, R. J. Zhang, H. J. Lai, and J. Zhang. 2017. “In-
MICP on engineering properties of cemented soils.” Géotechnique fluence of cementation level on the strength behaviour of bio-cemented
63 (4): 331–339. https://doi.org/10.1680/geot.SIP13.P.022. sand.” Acta Geotech. 12 (5): 971–986. https://doi.org/10.1007/s11440
Al Qabany, A., K. Soga, and C. Santamarina. 2012. “Factors affecting -017-0574-9.
efficiency of microbially induced calcite precipitation.” J. Geotech. DeJong, J. T., M. B. Fritzges, and K. Nüsslein. 2006. “Microbially induced
Geoenviron. Eng. 138 (8): 992–1001. https://doi.org/10.1061/(ASCE) cementation to control sand response to undrained shear.” J. Geotech.
GT.1943-5606.0000666. Geoenviron. Eng. 132 (11): 1381–1392. https://doi.org/10.1061/(ASCE)
ASTM. 2006. Standard test method for permeability of granular soils (con- 1090-0241(2006)132:11(1381).
stant head). ASTM D2434–68. West Conshohocken, PA: ASTM. Dick, J., W. De Windt, B. De Graef, H. Saveyn, P. Van Der Meeren,
ASTM. 2010. Standard practice for classification of soils for engineering N. De Belie, and W. Verstraete. 2006. “Bio-deposition of a calcium
purposes (unified soil classification system). ASTM D2487. West carbonate layer on degraded limestone by Bacillus species.” Biodegra-
Conshohocken, PA: ASTM. dation 17 (4): 357–367. https://doi.org/10.1007/s10532-005-9006-x.
ASTM. 2014. Standard specification for sand. ASTM C778. West Dilrukshi, R. A. N., K. Nakashima, and S. Kawasaki. 2018. “Soil improve-
Conshohocken, PA: ASTM. ment using plant-derived urease-induced calcium carbonate precipita-
Bernardi, D., J. T. Dejong, B. M. Montoya, and B. C. Martinez. 2014. tion.” Soils Found. 58 (4): 894–910. https://doi.org/10.1016/j.sandf
“Bio-bricks: Biologically cemented sandstone bricks.” Constr. Build. .2018.04.003.
Mater. 55 (Mar): 462–469. https://doi.org/10.1016/j.conbuildmat.2014 Dilrukshi, R. A. N., J. Watanabe, and S. Kawasaki. 2016. “Strengthening of
.01.019. sand cemented with calcium phosphate compounds using plant-derived
Burbank, M. B., T. J. Weaver, T. L. Green, B. C. Williams, and R. L. urease.” Int. J. GEOMATE 11 (25): 2461–2467. https://doi.org/10
Crawford. 2011. “Precipitation of calcite by indigenous microorgan- .21660/2016.25.5149.
isms to strengthen liquefiable soils.” Geomicrobiol. J. 28 (4): 301–312.
Faibish, R. S., M. Elimelech, and Y. Cohen. 1998. “Effect of interparticle
https://doi.org/10.1080/01490451.2010.499929.
electrostatic double layer interactions on permeate flux decline in cross-
Carmona, J. P. S. F., P. J. V. Oliveira, L. J. L. Lemos, and A. M. G. Pedro.
flow membrane filtration of colloidal suspensions: An experimental in-
2018. “Improvement of a sandy soil by enzymatic calcium carbonate
vestigation.” J. Colloid Interface Sci. 204 (1): 77–86. https://doi.org/10
precipitation.” Proc. Inst. Civ. Eng. Geotech. Eng. 171 (1): 3–15.
.1006/jcis.1998.5563.
https://doi.org/10.1680/jgeen.16.00138.
Feng, K., and B. M. Montoya. 2015. “Influence of confinement and cemen-
Cheng, L., R. Cord-Ruwisch, and M. A. Shahin. 2013. “Cementation of
tation level on the behavior of microbial-induced calcite precipitated
sand soil by microbially induced calcite precipitation at various degrees
of saturation.” Can. Geotech. J. 50 (1): 81–90. https://doi.org/10.1139 sands under monotonic drained loading.” J. Geotech. Geoenviron. Eng.
/cgj-2012-0023. 142 (1): 04015057. https://doi.org/10.1061/(ASCE)GT.1943-5606
Cheng, L., M. A. Shahin, and D. Mujah. 2017. “Influence of key environ- .0001379.
mental conditions on microbially induced cementation for soil stabili- Foppen, J. W. A., and J. F. Schijven. 2006. “Evaluation of data from the
zation.” J. Geotech. Geoenviron. Eng. 143 (1): 04016083. https://doi literature on the transport and survival of Escherichia coli and thermo-
.org/10.1061/(ASCE)GT.1943-5606.0001586. tolerant coliforms in aquifers under saturated conditions.” Water Res.
Choi, S., T. Hoang, E. J. Alleman, and J. Chu. 2019a. “Splitting tensile 40 (3): 401–426. https://doi.org/10.1016/j.watres.2005.11.018.
strength of fiber-reinforced and biocemented sand.” J. Mater. Civ. Eng. Ginn, T. R., B. D. Wood, K. E. Nelson, T. D. Scheibe, E. M. Murphy, and
31 (9): 1–6. https://doi.org/10.1061/(ASCE)MT.1943-5533.0002841. T. P. Clement. 2002. “Processes in microbial transport in the natural
Choi, S., T. Hoang, and S. Park. 2019b. “Undrained behavior of microbially subsurface.” Adv. Water Resour. 25 (8–12): 1017–1042. https://doi
induced calcite precipitated sand with polyvinyl alcohol fiber.” Appl. .org/10.1016/S0309-1708(02)00046-5.
Sci. 9 (6): 1214. https://doi.org/10.3390/app9061214. Gomez, M. G., B. C. Martinez, and J. T. Dejong. 2013. “Bio-mediated soil
Choi, S., S. Wu, and J. Chu. 2016a. “Biocementation for sand using an improvement field study to stabilize mine sands.” In Proc., GeoMontreal.
eggshell as calcium source.” J. Geotech. Geoenviron. Eng. 142 (10): Montreal: Ontario Water Works Association.
06016010. https://doi.org/10.1061/(ASCE)GT.1943-5606.0001534. Gomez, M. G., B. C. Martinez, J. T. DeJong, C. E. Hunt, L. A. deVlaming,
Choi, S. G., J. Chu, R. C. Brown, K. Wang, and Z. Wen. 2017a. “Sustain- D. W. Major, and S. M. Dworatzek. 2015. “Field-scale bio-cementation
able biocement production via microbially induced calcium carbonate tests to improve sands.” Proc. Inst. Civ. Eng. Ground Improv. 168 (3):
precipitation: Use of limestone and acetic acid derived from pyrolysis 206–216. https://doi.org/10.1680/grim.13.00052.

© ASCE 04020030-13 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2020, 32(4): 04020030


Hamdan, N., and E. Kavazanjian. 2016. “Enzyme-induced carbonate its effects on strength of coarse and fine grained sand.” Ecol. Eng.
mineral precipitation for fugitive dust control.” Géotechnique 66 (7): 109 (Sep): 57–64. https://doi.org/10.1016/j.ecoleng.2017.09.011.
546–555. https://doi.org/10.1680/jgeot.15.P.168. Nafisi, A., S. Safavizadeh, and B. M. Montoya. 2019. “Influence of microbe
Hamdan, N., E. Kavazanjian Jr., and S. O’Donnell. 2013. “Carbonate and enzyme-induced treatments on cemented sand shear response.”
cementation via plant derived urease.” In Proc., 18th Int. Conf. on Soil J. Goetech. Geoenviron. Eng. 145 (9): 06019008. https://doi.org/10
Mechanics and Geotechnical Engineering, 2489–2492. Paris: French .1061/(ASCE)GT.1943-5606.0002111.
Society for Soil Mechanics and Geotechnical Engineering. Nam, I., C. Chon, K. Jung, and S. Choi. 2015. “Calcite precipitation
Hamed Khodadadi, T., E. Kavazanjian, L. Van Paassen, and J. Dejong. by ureolytic plant (Canavalia ensiformis) extracts as effective bioma-
2017. “Bio-grout materials: A review.” In Proc., Grouting 2017. terials.” KSCE J. Civ. Eng. 19 (6): 1620–1625. https://doi.org/10.1007
Reston, VA: ASCE. /s12205-014-0558-3.
Hoang, T., J. Alleman, B. Cetin, K. Ikuma, and S.-G. Choi. 2018. “Sand Nemati, M., and G. Voordouw. 2003. “Modification of porous media
Downloaded from ascelibrary.org by Nanyang Technological University- Library on 02/04/20. Copyright ASCE. For personal use only; all rights reserved.

and silty-sand soil stabilization using bacterial enzyme induced calcite permeability, using calcium carbonate produced enzymatically in situ.”
precipitation (BEICP).” Can. Geotech. J. 56 (6): 808–822. https://doi Enzyme Microb. Technol. 33 (5): 635–642. https://doi.org/10.1016
.org/10.1139/cgj-2018-0191. /S0141-0229(03)00191-1.
Ivanov, V., and J. Chu. 2008. “Applications of microorganisms to geotech- Neupane, D., H. Yasuhara, N. Kinoshita, and Y. Ando. 2015a. “Distribution
nical engineering for bioclogging and biocementation of soil in situ.” of mineralized carbonate and its quantification method in enzyme medi-
Rev. Environ. Sci. Biotechnol. 7 (2): 139–153. https://doi.org/10.1007 ated calcite precipitation technique.” Soils Found. 55 (2): 447–457.
/s11157-007-9126-3. https://doi.org/10.1016/j.sandf.2015.02.018.
Javadi, N., H. Khodadadi, N. Hamdan, and E. Kavazanjian Jr. 2018. “EICP Neupane, D., H. Yasuhara, N. Kinoshita, and H. Putra. 2015b. “Distribution
treatment of soil by using urease enzyme extracted from watermelon of grout material within 1-m sand column in insitu calcite precipitation
seeds.” In Proc., IFCEE 2018, 115–124. Reston, VA: ASCE. technique.” Soils Found. 55 (6): 1512–1518. https://doi.org/10.1016/j
Kaur, N., M. S. Reddy, and A. Mukherjee. 2016. “Significant indicators .sandf.2015.10.015.
for biomineralisation in sand of varying grain sizes.” Constr. Build. Neupane, D., H. Yasuhara, N. Kinoshita, and T. Unno. 2013. “Applicability
Mater. 104 (Feb): 198–207. https://doi.org/10.1016/j.conbuildmat.2015 of enzymatic calcium carbonate precipitation as a soil-strengthening
.12.023. technique.” J. Geotech. Geoenviron. Eng. 139 (12): 2201–2211. https://
Kavazanjian, E., Jr., A. Almajed, and N. Hamdan. 2017. “Bio-inspired soil doi.org/10.1061/(ASCE)GT.1943-5606.0000959.
improvement using EICP soil columns and soil nails.” In Proc., Grout- Oliveira, P. J. V., L. D. Freitas, and J. P. S. F. Carmona. 2016. “Effect of soil
ing 2017 GSP, 13–22. Reston, VA: ASCE. type on the enzymatic calcium carbonate precipitation process used for
Kavazanjian, E., and N. Hamdan. 2015. “Enzyme induced carbonate pre- soil improvement.” J. Mater. Civ. Eng. 29 (4): 04016263. https://doi.org
cipitation (EICP) columns for ground improvement.” In Proc., IFCEE /10.1061/(ASCE)MT.1943-5533.0001804.
2015, edited by M. Iskander, M. T. Suleiman, J. B. Anderson, and D. F. Oliveira, P. J. V., and J. P. G. Neves. 2019. “Effect of organic matter content
Laefer, 2252–2261. Reston, VA: ASCE. on enzymatic biocementation process applied to coarse-grained soils.”
Larsen, J., M. Poulsen, T. Lundgaard, and M. Agerbaek. 2008. “Plugging J. Mater. Civ. Eng. 31 (7): 04019121. https://doi.org/10.1061/(ASCE)
of fractures in chalk reservoirs by enzymatic-induced calcium carbonate MT.1943-5533.0002774.
precipitation.” SPE Production Oper. 23 (4): 478–483. https://doi.org Park, S.-S., S.-G. Choi, and I.-H. Nam. 2014. “Effect of microbially in-
/10.2118/108589-PA. duced calcite precipitation on strength of cemented sand.” J. Mater.
Lin, H., M. T. Suleiman, D. G. Brown, and E. Kavazanjian. 2016. Civ. Eng. 26 (8): 06014017. https://doi.org/10.1061/(ASCE)MT.1943
“Mechanical behavior of sands treated by microbially induced carbon- -5533.0001029.
ate precipitation.” J. Geotech. Geoenviron. Eng. 142 (2): 04015066. Pasillas, J. N., H. Khodadadi, K. Martin, P. Bandini, C. M. Newtson, and
https://doi.org/10.1061/(ASCE)GT.1943-5606.0001383. E. Kavazanjian Jr. 2018. “Viscosity-enhanced EICP treatment of soil.”
Martinez, B. C., and J. T. DeJong. 2009. “Bio-mediated soil improvement: In Proc., IFCEE 2018, 145–154. Reston, VA: ASCE.
Load transfer mechanisms at the micro- and macro- scales.” In Proc., Phua, Y. J., and A. Røyne. 2018. “Bio-cementation through controlled
US-China Workshop on Ground Improvement Technologies 2009, dissolution and recrystallization of calcium carbonate.” Constr. Build.
242–251. Reston, VA: ASCE. Mater. 167 (Apr): 657–668. https://doi.org/10.1016/j.conbuildmat.2018
Martinez, B. C., J. T. DeJong, T. R. Ginn, B. M. Montoya, T. H. Barkouki, .02.059.
C. Hunt, B. Tanyu, and D. Major. 2013. “Experimental optimization Putra, H., H. Yasuhara, and N. Kinoshita. 2017. “Applicability of nat-
of microbial-induced carbonate precipitation for soil improvement.” ural zeolite for NH-Forms removal in enzyme-mediated calcite pre-
J. Geotech. Geoenviron. Eng. 139 (Apr): 587–598. https://doi.org/10 cipitation technique.” Geosciences 7 (3): 61. https://doi.org/10.3390
.1061/(ASCE)GT.1943-5606.0000787. /geosciences7030061.
Mitchell, A. C., and F. G. Ferris. 2005. “The coprecipitation of Sr into cal- Rebata-Landa, V., and J. C. Santamarina. 2006. “Mechanical limits to
cite precipitates induced by bacterial ureolysis in artificial groundwater: microbial activity in deep sediments.” Geochem. Geophys. Geosyst.
Temperature and kinetic dependence.” Geochim. Cosmochim. Acta 7 (11): 1–12. https://doi.org/10.1029/2006GC001355.
69 (17): 4199–4210. https://doi.org/10.1016/j.gca.2005.03.014. Saneiyan, S., D. Ntarlagiannis, J. Ohan, J. Lee, F. Colwell, and S. Burns.
Mitchell, A. C., and F. G. Ferris. 2006. “The influence of Bacillus pasteurii 2019. “Induced polarization as a monitoring tool for in-situ micro-
on the nucleation and growth of calcium carbonate.” Geomicrobiol. J. bial induced carbonate precipitation (MICP) processes.” Ecol. Eng.
23 (3–4): 213–226. https://doi.org/10.1080/01490450600724233. 127 (Feb): 36–47. https://doi.org/10.1016/j.ecoleng.2018.11.010.
Mitchell, J. K. 1981. “Soil improvement state of the art report.” In Vol. 4 of Simatupang, M., and M. Okamura. 2017. “Liquefaction resistance of sand
Proc., 11th Int. Conf. on SMFE. Boca Raton, FL: CRC Press. remediated with carbonate precipitation at different degrees of satura-
Mortensen, B. M., M. J. Haber, J. T. Dejong, L. F. Caslake, and D. C. tion during curing.” Soils Found. 57 (4), 619–631. https://doi.org/10
Nelson. 2011. “Effects of environmental factors on microbial induced .1016/j.sandf.2017.04.003.
calcium carbonate precipitation.” J. Appl. Microbiol. 111 (2): 338–349. Sondi, I., and B. Salopek-Sondi. 2005. “Influence of the primary structure
https://doi.org/10.1111/j.1365-2672.2011.05065.x. of enzymes on the formation of CaCO3 polymorphs: A comparison of
Mujah, D., L. Cheng, and M. A. Shahin. 2019. “Microstructural and plant (Canavaliaensiformis) and bacterial (Bacilluspasteurii) ureases.”
geomechanical study on biocemented sand for optimization of MICP Langmuir 21 (19): 8876–8882. https://doi.org/10.1021/la051129v.
process.” J. Mater. Civ. Eng. 31 (4): 04019025. https://doi.org/10.1061 Stocks-Fischer, S., J. K. Galinat, and S. S. Bang. 1999. “Microbiological
/(ASCE)MT.1943-5533.0002660. precipitation of CaCO3 .” Soil Biol. Biochem. 31 (11): 1563–1571.
Mwandira, W., K. Nakashima, and S. Kawasaki. 2017. “Bioremediation https://doi.org/10.1016/S0038-0717(99)00082-6.
of lead-contaminated mine waste by Pararhodobacter sp. based on Sun, X., L. Miao, D. Ph, T. Tong, and C. Wang. 2018. “Improvement of
the microbially induced calcium carbonate precipitation technique and microbial-induced calcium carbonate precipitation technology for sand

© ASCE 04020030-14 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2020, 32(4): 04020030


solidification.” J. Mater. Civ. Eng. 30 (11): 04018301. https://doi.org van Paassen, L. A., M. C. M. Van Loosdrecht, M. Pieron, A. Mulder,
/10.1061/(ASCE)MT.1943-5533.0002507. D. J. M. Ngan-Tillard, and T. J. M. Van Der Linden. 2010b. “Strength
Terashi, M., and I. Juran. 2000. “Ground improvement—State of the art.” and deformation of biologically cemented sandstone.” In Proc., Rock
In Proc., ISRM Int. Symp. Lisbon, Portugal: International Society for Engineering in Difficult Ground Conditions—Soft Rocks and Karst,
Rock Mechanics and Rock Engineering. edited by I. Vrkljan, 405–410. Boca Raton, FL: CRC Press.
Terzis, D., and L. Laloui. 2018. “3-D micro-architecture and mechanical Weil, M. H., J. T. Dejong, B. C. Martinez, and B. M. Mortensen. 2018.
response of soil cemented via microbial-induced calcite precipitation.” “Seismic and resistivity measurements for real-time monitoring of mi-
Sci. Rep. 8 (1): 1416. https://doi.org/10.1038/s41598-018-19895-w. crobially induced calcite precipitation in sand.” Geotech. Test. J. 35 (2):
Terzis, D., and L. Laloui. 2019. “Cell-free soil bio-cementation with 330–341. https://doi.org/10.1520/GTJ103365.
strength, dilatancy and fabric characterization.” Acta Geotech. 14 (3): Whiffin, V. S. 2004. “Microbial CaCO3 precipitation for the production of
639–656. https://doi.org/10.1007/s11440-019-00764-3.
biocement.” Ph.D. dissertation, School of Biological Sciences and Bio-
van der Star, W. R. L., W. K. van Wijngaarden-van Rossum, L. A. van
Downloaded from ascelibrary.org by Nanyang Technological University- Library on 02/04/20. Copyright ASCE. For personal use only; all rights reserved.

technology, Murdoch Univ.


Paassen, and L. R. van Baalen. 2011. “Stabilization of gravel deposits
Whiffin, V. S., L. A. van Paassen, and M. P. Harkes. 2007. “Microbial car-
using microorganisms.” In Proc., 15th European Conf. on Soil Mechan-
bonate precipitation as a soil improvement technique.” Geomicrobiol. J.
ics and Geotechnical Engineering, edited by A. Anagnostopoulos,
85–90. Amsterdam, Netherlands: IOS Press. 24 (5): 417–423. https://doi.org/10.1080/01490450701436505.
van Paassen, L. A. 2011. “Bio-mediated ground improvement: From Yasuhara, H., D. Neupane, K. Hayashi, and M. Okamura. 2012. “Experi-
laboratory experiment to pilot applications.” In Proc., Geo-Frontiers, ments and predictions of physical properties of sand cemented by
4099–4108. Reston, VA: ASCE. enzymatically-induced carbonate precipitation.” Soils Found. 52 (3):
van Paassen, L. A., R. Ghose, T. J. M. van der Linden, W. R. L. van der Star, 539–549. https://doi.org/10.1016/j.sandf.2012.05.011.
and M. C. M. van Loosdrecht. 2010a. “Quantifying biomediated ground Zhao, Q., L. Li, C. Li, M. Li, F. Amini, and H. Zhang. 2014. “Factors
improvement by ureolysis: Large-scale biogrout experiment.” J. Geo- affecting improvement of engineering properties of MICP-treated
tech. Geoenviron. Eng. 136 (12): 1721–1728. https://doi.org/10.1061 soil catalyzed by bacteria and urease.” J. Mater. Civ. Eng. 26 (12):
/(ASCE)GT.1943-5606.0000382. 04014094. https://doi.org/10.1061/(ASCE)MT.1943-5533.0001013.

© ASCE 04020030-15 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2020, 32(4): 04020030

You might also like