You are on page 1of 12

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/322537057

Liquefaction resistance of bio-cemented calcareous sand

Article  in  Soil Dynamics and Earthquake Engineering · April 2018


DOI: 10.1016/j.soildyn.2018.01.008

CITATIONS READS

13 883

5 authors, including:

Peng Xiao Hong Liu


Chongqing University Chongqing University
7 PUBLICATIONS   36 CITATIONS    155 PUBLICATIONS   1,484 CITATIONS   

SEE PROFILE SEE PROFILE

Yang Xiao Armin Stuedlein


Chongqing University Oregon State University
94 PUBLICATIONS   947 CITATIONS    90 PUBLICATIONS   538 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Critical state behaviors and bounding surface model of rockfill material considering particle breakage View project

High pressure jet grouting technqiue View project

All content following this page was uploaded by Yang Xiao on 17 January 2018.

The user has requested enhancement of the downloaded file.


Soil Dynamics and Earthquake Engineering 107 (2018) 9–19

Contents lists available at ScienceDirect

Soil Dynamics and Earthquake Engineering


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

Liquefaction resistance of bio-cemented calcareous sand T


a,b a,b a,b,⁎ c c
Peng Xiao , Hanlong Liu , Yang Xiao , Armin W. Stuedlein , T. Matthew Evans
a
School of Civil Engineering, Chongqing University, Chongqing, China
b
Key Laboratory of New Technology for Construction of Cities in Mountain Area, Chongqing University, Chongqing 400045, China
c
School of Civil and Construction Engineering, Oregon State University, Corvallis, OR, USA

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

Keywords: Coral reefs and other calcareous deposits may experience various types of significant dynamic loading, such as
Calcareous sand those from waves and earthquakes. When submerged and subjected to earthquake loading, the potential for
Liquefaction mitigation liquefaction of calcareous deposits may cause a loss of human life and property; however, few studies have
Microbial induced calcite precipitation evaluated the liquefaction potential of calcareous sands relative to those conducted on silica sands. Accordingly,
Cyclic testing
it is critical to study the cyclic resistance of calcareous sands as well as methods to mitigate their liquefaction
Cyclic strength
potential. Microbial induced calcite precipitation (MICP) offers one such strategy that can be considered for
improving the cyclic resistance of calcareous sands, particularly for those applications below existing infra-
structure that would pose technical difficulties for traditional modes of ground improvement. This paper ex-
amines the effectiveness of MICP on the cyclic resistance of as a function of cementation solution (CS) content,
effective confining pressure, and cyclic stress ratio (CSR) through a cyclic triaxial test program. The generation
and accumulation of excess pore pressure and corresponding axial strains are compared across a range of treated
and untreated sands. This study shows that the liquefaction resistance of clean calcareous sand may be sig-
nificantly improved by the MICP treatment. Scanning electron microscope images are presented to help link the
improvement in cyclic response to the microstructural features of the microbial-induced calcite and bio-ce-
mented sand.

1. Introduction result of various physical, mechanical, chemical and biological de-


positional environments [7]. Calcareous sands consist of shells and
Coral reefs and other calcareous soil deposits may experience var- corals, and as a result of their previous biological life, the sand include
ious types of significant dynamic loading, such as those from waves and cavities and voids, which can be easily crushed [8–11]. Due to their low
earthquakes. Additionally, significant coastal and ocean infrastructure mineral strength, it is necessary to improve the cyclic resistance of loose
are being constructed on these deposits [1,2]. Since 1960s, calcareous to medium dense deposits of calcareous sand supporting structures.
soil deposits have attracted increased attention for the exploitation of However, many traditional technologies used to improve liquefaction
hydrocarbon resources [3]. In ocean engineering, calcareous sands are resistance, such as densification through vibro-compaction and vibro-
commonly used as the foundation materials for buildings and break- replacement [12], driving of displacement piling [13,14], or compac-
waters and as the backfill material for road embankments or airport tion grouting [15], are inappropriate for improvement of calcareous
runways [4]. When submerged and subject to earthquake loading, the sands as these methods induce significant forces that can crush the sand
potential for liquefaction of calcareous deposits may cause a loss of grains [16–18]. On the other hand, other grouting techniques with
human life and property, as observed during the 1993 Guam earth- cement [19] or other cement agents [20], may be disqualified in certain
quake, which was characterized with a moment magnitude, Mw, of 7.7 marine settings due to environmental restrictions on such techniques
[5]. This earthquake caused significant liquefaction and resulting da- and might pollute the marine environment [21].
mage to the commercial facilities and port areas on the deposits, which Recently, a novel ground improvement technique [22–25] has
highlight the importance of understanding the liquefaction of calcar- gained interest due to its relatively green and sustainable advantages:
eous sand. Calcareous sands are broadly-distributed along the coasts microbial induced calcite precipitation (MICP). Ivanov and Chu [24]
and coastal reefs of Australia, the Persian Gulf, the Gulf of Mexico, and suggested that the raw materials for microbial grouting are significantly
the South China Sea for example [6], and these types of soils are the cheaper than chemical grout. The basic mechanisms of improvement


Corresponding author at: School of Civil Engineering, Chongqing University, Chongqing, China.
E-mail address: hhuxyanson@163.com (Y. Xiao).

https://doi.org/10.1016/j.soildyn.2018.01.008
Received 9 September 2017; Received in revised form 15 November 2017; Accepted 2 January 2018
0267-7261/ © 2018 Elsevier Ltd. All rights reserved.
P. Xiao et al. Soil Dynamics and Earthquake Engineering 107 (2018) 9–19

using MICP technique includes: (1) synthesis of an enzyme through


bacteria metabolic activities [26]; (2) microbial urease hydrolyses urea
to produce ammonia and carbonate ions; (3) the carbonate ions bind
with calcium to accumulate insoluble CaCO3 in a calcium rich en-
vironment [27]. The chemical reactions governing the MICP process are
given by:
CO(NH2)2 +2H2 O →→ 2NH+4 + CO32 − (1)

Ca2 + + CO32 − →→ CaCO3 (2)


Bio-cementation has been recognized for its broad application in
treating various engineering challenges. For example, MICP has been
used to reduce the permeability of soil [28,29], repair cracks in con-
crete [30,31], and improve the stiffness and strength of soil [32–36].
Additionally, several studies have shown that the bio-cementation
technology has significant potential to efficiently mitigate liquefaction
at developed sites in a passive, non-disruptive manner [37–40]. Bur- Fig. 1. Particle size distribution curve superimposed with liquefiable boundary (modified
from Diaz-Rodriguez et al., 2008).
bank [38] et al. compared cyclic shear test results of sand treated with
bio-cementation and Portland cement, and found the cyclic stress ratio
(CSR) of the bio-treated soil was higher than the cement-treated soil for Table 1
similar quantities of cementing agent. Han et al. [37] found that Index properties for the original calcareous sand.

treating sand with MICP takes less time to improving liquefiable sand
Sand property Values
than colloidal silica grouting to achieve a similar liquefaction re-
sistance. Montoya et al. [40] studied the dynamic response of MICP- Specific Gravity (Gs) 2.79
treated siliceous sand of various strengths using centrifuge models and Minimun Void Ratio (emin) 1.22
Maximum Void Ratio (emax) 1.69
found that the pore pressure development, surface settlement, and the
D10 0.18
acceleration at the ground surface was lower than that of a comparable D30 0.28
loose sand. On the other hand, Sasaki and Kuwano [41] found that si- D50 0.36
liceous sand with a 30% clay fraction showed no increase in liquefac- D60 0.40
tion resistance following microbial precipitation of CaCO3. Coefficient of Uniformity (Cu) 2.22
Coefficient of Curvature (Cc) 1.09
There are far fewer studies that have evaluated how the magnitude
of MICP improvement and confining pressure has impacted liquefaction
resistance. In addition, the few studies that have focused on the cyclic other carbonate sands [7,42,43,46]. Based on the Unified Soil Classi-
response of marine soils like calcareous sand [42,43] have evaluated fication System [47], the sand is classified as poorly-graded, other
natural sands without MICP treatment. Calcareous sand is a biogenic properties such as the specific gravity, and maximum and minimum
material that shares many physico-chemical characteristics with cal- void ratios are shown in Table 1. The CaCO3 content in Yongxing Island
cium-based bio-cementation; accordingly, the potential for natural ef- sand was more than 90% prior to MICP treatment. Two sets of particle
ficiencies between the carbonate crystal-particle interface strength may size curves including the ranges in PSDs for most liquefiable and po-
be significant. For example, Khan et al. [44,45] conducted some initial tentially liquefiable soils [48] are compared to Yongxing Island sand in
investigations on the coral sand using different bacteria and observed Fig. 1. Based on the PSD, the potential for liquefaction of loose deposits
that unconfined compressive strengths in the range of 13–20 MPa were of this sand is high.
achievable.
The objective of this experimental laboratory investigation is to
determine the efficacy of MICP-treatment to improve the cyclic re- 2.2. Bacterial suspension and cementation solution (CS)
sistance of calcareous sand. In order to make sufficient comparison
against various treatments, the cyclic response of the untreated cal- The urease strain of Sporosarcina pasteurii was used as the active
careous sand is studied using a series of cyclic triaxial compression tests microbe for production of the calcite in the experiments conducted in
(CTCs). Then, a series of CTCs are conducted to compare the liquefac- this study. The previously frozen strain was activated in a plate
tion mitigation of different cementation solution (CS) contents, effec- medium, and then cultivated in sterilized growth media which con-
tive confining pressures, and cyclic stress ratios and to analyze the sisted of the following per-liter deionized water concentrations of:(1)
factors affecting mitigation. Scanning electron microscope (SEM) 20 g yeast extract, (2) 10 g NH4Cl, (3) 12 mg MnCl2·H20, (4) 24 mg
images are presented to help link the improvement in cyclic response to NiCl2·6H20 with a pH of 9.0, achieved using sodium hydroxide. Bacteria
the microstructural features of the microbial-induced calcite and bio- were grown in incubator shaker in 30 °C with about 36 h to reach an
cemented sand. optical density of 600 nm (OD600) of 0.8–1.0 (107 cells/mL). To remove
the abundant metabolic waste and odor, the harvested bacteria were
2. Experimental materials centrifuged at 4 °C at a speed of 4000 rpm for 15 min. Following con-
centration in the centrifuge, the resulting supernatant was removed and
2.1. Characteristics of Yongxing Sand replaced with 0.9% sodium chloride solution before the bacteria were
resuspended. The enzymatic activity of the final bacterial media was
The calcareous sands in this study were sourced from Yongxing about 1.4–2.0 mM urea/min. The cementation solution (CS) used in the
Island of the Xisha archipelago located in the South China Sea. The experiments was a mixture of 0.5 mol/L urea and 0.5 mol/L CaCl2.
particle size distribution (PSD) and a microscopic image of the
Yongxing Island sands are presented in Fig. 1. The medium to coarse,
shelly sand is characterized with widely varying grain shapes ranging
from platy to rod-like to ellipsoidal; these grain shapes can facilitate
significant interlock while exhibiting remarkably large void space like

10
P. Xiao et al. Soil Dynamics and Earthquake Engineering 107 (2018) 9–19

Fig. 2. Test setup schematics: (a) Picture of MICP treatment procedure; (b) Sketch of MICP treatment procedure. (a. Cementation solution/ Bacterial suspension; b. Peristaltic pump; c.
PVC split mold; d. Scouring pad; e. Tightening ring; f. Aspirating hole; g. Rubber membrane; h. Sand column; i. Rubber plug; j. Drainage valve; k. Effluent).

3. Experimental procedures adjusted to maintain a small head of CS at the top of the specimen.
When a given volume of CS had been pumped into the soil column, the
3.1. Sample preparation drainage valve was closed and 6 h of reaction time was allowed to
elapse, completing one CS cycle.
Several previous studies on the static triaxial response MICP-treated Additional cycles were conducted in order to complete the Ca2+
soils were conducted by precipitating the bio-cement directly in the reaction, which was determined by the measurement of pH and Ca2+
triaxial cell and apparatus [32,35]. In this study, specimens were pro- concentration in the effluent. The target volume of CS for different
duced by using a set of specially-designed specimen preparation devices treatment groups varied as shown in Table 2, and the measured change
to achieve batches of specimens. The procedure of preparing specimens in mass (Δm) of the oven-dried, treated sand was used to indicate the
is shown in Fig. 2. Fig. 2(a) depicts the MICP treatment procedure with global uniformity in treatment within a given test group.
eight soil columns being treated per pump, whereas Fig. 2(b) presents Following each treatment cycle, the Ca2+ concentration was mea-
the procedure schematically. As shown in Fig. 2(b), the test apparatus sured using the EDTA titration method described by Achal et al. [49].
consists of a polyvinyl chloride (PVC) split mold, rubber plug, rubber Upon completion of the reaction and following a rest period of ap-
membrane, scouring pad, and tightening ring. The PVC split mold has proximately 24 h, 300 mL of deionized water was circulated through
an internal diameter and height of 4 and 12.5 cm, respectively, and is the specimen to remove remaining nutrients. To facilitate comparison
transparent to allow visual inspection of the treatment process. to the treated specimens, the untreated clean sand specimens were
Specimens were prepared by air-pluviation into a split-mold lined prepared using the same general procedures excluding the treatment
with a stretched membrane under vacuum. This technique proved to be process. However, specimen U12 was prepared as a control specimen to
advantageous over other sample preparation techniques as it eliminated specifically to measure the changes in pH and Ca2+ concentration that
disturbance due to extrusion from the treatment column. The test sand could develop for numerous cycles of CS in the absence of bacterial
were went across a No. 10 (i.e., 2 mm) sieve to remove organic debris suspension, as described below.
(e.g., plant matter) and provide uniformity in the maximum grain size.
Then the sand were over-dried at 60 °C (the particle surface of calcar-
eous sand would get bark when over-dried over 100 °C). Following 3.3. Undrained cyclic triaxial tests
assembly of the treatment column and treatment (described below),
specimens of about 8 cm height and 3.9 cm diameter were produced for Undrained cyclic triaxial (UCT) tests were conducted on the pre-
cyclic triaxial testing. The air-pluviation procedures were selected to pared specimens of clean and MICP-treated calcareous sand. Prior to
produce a medium dense condition after pluviation (Table 2), and the removal of the split-mold, the top and bottom surface of each specimen
mean and coefficient of variation for the relative density of the speci- was leveled and the specimen placed on a filter paper-lined porous
mens were 47% and 5%, respectively. stone located at the base of the triaxial apparatus pedestal. Then the top
and bottom of the membrane was rolled up and down to cover the top
cap and pedestal and sealed with O-rings. A small vacuum pressure of
3.2. MICP treatment procedure
less than 20 kPa was maintained in order to preserve the integrity of the
specimen. The tests were carried out in general accordance with ASTM
Upon assembly of the sand column apparatus and specimen plu-
D5311 [50]. To aid saturation, samples were first flushed with CO2 and
viation, the specimen was flushed with deionized water and the drai-
de-aired water under an effective stress of 20 kPa, then backpressure
nage outlet was closed and raised up to the same level of the sample
saturation of 300 kPa was conducted in order to obtain a B-value of
surface to achieve equality of total head. Thereafter, the bacterial sus-
0.95 or greater. Specimens were loaded cyclically using sinusoidal de-
pension was pumped through the specimen at a rate of 80 mL/h to
viatoric stresses at a frequency of 1 Hz. Liquefaction (NL) is defined in
distribute the suspension throughout the pore volume. A detention time
this study as excess pore water pressure ratio (ru=Δu/σ′c ) greater than
of 6 h was provided to allow the bacteria to disperse and adhere to the
0.95. During these tests, axial stress, axial displacement, and pore water
soil grains.
pressure were monitored. Thirty-one UCTs were conducted, as shown in
Following detention of the bacterial suspension, several cycles of CS
Table 2, which summarizes the test conditions, including the ranges in
were pumped through the specimen as follows. Cementation solution
sand treatment, effective confining pressures, and cyclic stress ratio
was pumped into the soil sample at a rate of 40 mL/h with the effluent
(CSRs) applied during the undrained cyclic shear.
drainage valve opened slightly less than the influent to promote storage
of CS within the pore volume. The drainage valve was continuously

11
P. Xiao et al. Soil Dynamics and Earthquake Engineering 107 (2018) 9–19

Table 2
Properties of tested soil and test parameters.

Test no. Cementation solution (L) ρd (g/cm3) e0 Δm (g) Dr (%) B-value σ ′c (kPa) CSR NL

Group U
U1 0 1.132 1.463 / 48 0.98 50 0.167 312
U2 0 1.132 1.463 / 48 0.95 50 0.208 88
U3 0 1.129 1.470 / 47 0.97 50 0.250 41
U4 0 1.137 1.453 / 50 0.96 50 0.292 20
U5 0 1.134 1.460 / 49 0.97 100 0.146 460
U6 0 1.120 1.490 / 42 0.97 100 0.167 217
U7 0 1.134 1.460 / 49 0.96 100 0.188 103
U8 0 1.128 1.474 / 46 0.97 100 0.250 31
U9 0 1.120 1.490 / 42 0.97 200 0.167 125
U10 0 1.122 1.487 / 43 0.95 200 0.208 40
U11 0 1.137 1.453 / 50 0.97 200 0.250 11
U12a 0.6 1.132 1.463 / 48 / / / /
Group TA
TA1 0.2 1.126 1.477 5.64 45 0.97 100 0.188 435
TA2 0.2 1.137 1.453 5.83 50 0.97 100 0.208 137
TA3 0.2 1.132 1.463 5.64 48 0.95 100 0.250 72
TA4 0.2 1.126 1.477 5.75 45 0.98 100 0.292 21
Group TB
TB1 0.4 1.123 1.484 11.59 44 0.97 50 0.250 135
TB2 0.4 1.129 1.470 12.11 47 0.96 50 0.292 75
TB3 0.4 1.132 1.463 11.90 48 0.97 50 0.333 49
TB4 0.4 1.128 1.474 12.11 46 0.97 50 0.375 23
TB5 0.4 1.137 1.453 11.72 50 0.95 100 0.208 388
TB6 0.4 1.128 1.474 11.75 46 0.97 100 0.250 118
TB7 0.4 1.138 1.452 11.59 51 0.97 100 0.292 54
TB8 0.4 1.132 1.463 12.28 48 0.97 100 0.333 22
TB9 0.4 1.132 1.463 11.85 48 0.96 200 0.208 331
TB10 0.4 1.123 1.484 11.91 44 0.97 200 0.250 78
TB11 0.4 1.123 1.484 12.05 44 0.97 200 0.292 36
TB12 0.4 1.132 1.463 12.03 48 0.95 200 0.333 18
Group TC
TC1 0.6 1.128 1.474 17.29 46 0.97 100 0.229 250
TC2 0.6 1.137 1.453 17.83 50 0.98 100 0.250 135
TC3 0.6 1.132 1.463 17.79 48 0.97 100 0.292 60
TC4 0.6 1.128 1.474 17.35 46 0.98 100 0.333 38

Note:.
ρd =Initial dry density; e0 =Void ratio; σ ′c =Initial effective cell pressure; CSR=Cyclic stress ratio; NL =The number of cycles to liquefaction.
a
U12 conducted to illustrate changes in chemical composition from the specimen preparation techniques, only.

3.4. Scanning Electron Microscopy (SEM) imaging which was not dosed with bacterial solution prior to CS treatment.
Fig. 3(a) shows that treatment with the CS within specimens dosed with
Microscopic imaging was conducted to link the experimental cyclic the bacterial suspension produced a marked increase in pH after the
response of MICP-treated sand to the physical soil fabric produced using first cycle as a result of ammonium ion production from urea hydro-
bio-cementation. Imaging was conducted using a Zeiss Auriga-SEM; a lysis. Thereafter, the pH tended to decrease slightly. As shown in the
relatively low accelerating voltage of 5 kV was selected to ensure high Fig. 3(a), the pH of the treated specimens remained greater than the
resolution (but relatively less bright) images. Prior to conducting the control specimen U12. This observation is generally consistent with the
SEM imaging, the samples were flushed with deionized water and dried observations reported by others [52,53], and indicated that pH can
for 24 h. Thereafter, microscopic specimens were produced by adhering indicate the activity of urease during MICP treatment.
a small amount of sand to the specimen plate and then coating with Similarly, the Ca2+ concentration was measured following each
gold under vacuum conditions. Described below, the microscopic in- cycle, and was used to gage the end of the calcite precipitation reaction,
vestigation facilitates the connection of the crystal formation and indicated by the near-zero magnitude of Ca2+ (n.b., Ca2+ cannot be
bonding between the sand particles and CaCO3 crystals to the cyclic fully-consumed in the chemical reaction due to the volatilization of
response. urea; see Eq. (1)). As described earlier, the change in mass of the treated
sand (Table 2) was used as an index of the global magnitude of CaCO3
content produced as an alternate approach for the dissolution approach
4. Test results and discussion [54]. However, in the current work, the dissolution approach cannot be
used as it would not distinguish between the natural sand CaCO3 and
4.1. Physico-Chemical Characterization of the treated and untreated the MICP-induced CaCO3. The coefficient of variation in Δm in Table 2
samples ranges from 1.6% to 1.8% for each group, which indicates a sufficient
level of repeatability in the treatment regime adopted.
In the process of MICP treatment, the pH of the aqueous solution
during precipitation can be used as a key indicator of CaCO3 pre-
cipitation. The pH increases during urea hydrolysis via the enzyme 4.2. Undrained cyclic resistance
urease; the local rise in pH often causes the bacteria to serve as nu-
cleation sites for crystallization [51]. Fig. 3 presents the variation in pH The effectiveness of MICP-treatment for improving the undrained
and Ca2+ concentrations of the effluent taken from specimens following cyclic resistance was evaluated by comparing the undrained cyclic re-
each CS treatment cycle and compares the response to specimen U12, sponse of the treated and clean sand specimens under stress-controlled

12
P. Xiao et al. Soil Dynamics and Earthquake Engineering 107 (2018) 9–19

Fig. 3. Variation of the pH and Ca2+ over time during the MICP treatment: (a) pH vs. treatment cycles; (b) Ca2+ vs. treatment cycles.

Fig. 4. The typical undrained cyclic response of clean calcareous sand sheared with σ′c = 100 kPa and CSR=0.250 (Specimen U8): (a) excess pore pressure response, (b) axial strain, (c)
deviatoric stress, and (d) cyclic stress path.

conditions. Cyclic stress ratios, defined as the ratio of the applied cyclic mean excess pore pressure (Um) that is non-recoverable from cycle to
shear stress and the effective confining pressure were specified for each cycle. A red line in Figs. 4(a) and 5(a) indicates the point at which
UCT test (Table 2) to draw general trends between the number of cycles initial liquefaction is attained. For the clean untreated sand, initial li-
to liquefaction, NL, and the magnitude of shear stresses. The response of quefaction occurred after 31 cycles (i.e., NL=31). In contrast, the
typical untreated and treated specimens (U8 and TC2, respectively) treated specimen did not liquefy until NL=135.
sheared in the same condition (σ′c = 100kPa and CSR=0.250) are pre- As shown in Fig. 4, the untreated specimen exhibited a contractive
sented in Figs. 4 and 5, respectively. The two Figures depict the de- response during the first quarter cycle of loading. The initial axial
velopment of excess pore water pressure ratio and axial strain (εa ) with strains were very small for the first several cycles and experienced a
the number of loading cycles, deviatoric stress-strain curve, and cyclic slow increase in the accumulation of strain. The axial strain was −2.1%
effective stress path. Figs. 4(a) and 5(a) show that the excess pore at NL=31, and increased to 4.6% by N=37. In contrast, the deforma-
pressure generated during cyclic loading can be separated into two tion characteristics of the MICP-treated specimen shown in Fig. 5(b)
components: (1) the component associated with the loading cycle (Ucyc) depart significantly from the untreated sand. For example, at N=31,
and is generated and dissipated during each loading cycle; and (b) the the axial strain was 0.1%, and achieved liquefaction at −1.8% strain at

13
P. Xiao et al. Soil Dynamics and Earthquake Engineering 107 (2018) 9–19

Fig. 5. The undrained cyclic response of MICP-treated calcareous sand sheared with σ′c = 100kPa and CSR=0.250 (Specimen TC2): (a) excess pore pressure response, (b) axial strain, (c)
deviatoric stress, and (d) cyclic effective stress path.

NL=135. Separately, it is observed that the accumulation of strains was


not symmetric, but rather accumulated in a manner biased towards
extension (Figs. 4b and c, 5b and c), consistent with the results reported
by others [7,42,43]. The deformation characteristics for different MICP-
treated specimens is explored in greater below.
The deviatoric stresses and effective stress paths of the untreated
and treated specimens are shown in Figs. 4(c-d) and 5(c-d), respec-
tively. As the number of cycles increases, the specimens lose stiffness
during the loading cycle prior to exhibiting dilation hardening, and the
effective stress paths move toward the left in response to a reduced
mean effective stress produced by the accumulation of excess pore
pressure. Upon reaching initial liquefaction, the effective stress path of
the untreated specimen intersected the effective failure envelope of the
material and the soil reaches a state of near-zero mean effective stress,
and the deviatoric shear stress could no longer be sustained by the
specimen during the majority of each loading cycle. Although the
treated specimen also exhibited a loss of initial stiffness and increase in Fig. 6. CSR versus NL for calcareous sand treated by different CS content under 100 kPa
effective confining pressure.
pore pressure, its cyclic resistance was significantly greater as shown by
the magnitude of NL and the steeper failure envelope following lique-
faction. The comparison of cyclic response in Figs. 4 and 5 shows that treated calcareous sand can be expressed by an empirical equation
MICP-treated calcareous sand exhibits a significantly higher cyclic [55,56] in the following from:
shear resistance due to increased dilatancy, with characteristics similar
CSR = a (NL)−b (3)
to dense sands, and associated with the particle-to-particle and particle-
to-calcite interface bond strength. where a and b are empirical parameters, determined using ordinary
Fig. 6 compares the trends of CSR with NL for the untreated and least squares (OLS) or other approaches. The factor a is equivalent to
MICP-treated specimens prepared with 0.2, 0.4, and 0.6 L CS tested at the CSR of the soil for one cycle of loading and varies with factors that
an effective confining pressure of σ′C = 100 kPa. The number of cycles affect the liquefaction resistance; factor b relates the CSR to the influ-
to liquefaction increases as the magnitude of CSR decreases. Fig. 6 ence of the amplitude of the shear stress cycles, confining pressures, and
shows that the CSR increases significantly with increasing CS content soil characteristics. For comparison of MICP-treated specimens in Fig. 6,
up to 0.4 L CS, after which the rate of increase in CSR reduces. the data are fit with CSR = a (NL)−b . The value of a for the specimens of
Trends in the cyclic failure curves for the untreated and MICP- MICP=0, 0.2 CS, 0.4 CS, 0.6 CS are 0.477, 0.522, 0.591 and 0.620

14
P. Xiao et al. Soil Dynamics and Earthquake Engineering 107 (2018) 9–19

Fig. 8. Development of excess pore pressure with number of cycles for untreated and
treated calcareous sand under 100 kPa effective confining pressure and CSR=0.25.

envelope associated with decreased resistance to liquefaction [61,62].

4.3. Development of excess pore pressure

Evaluation of the generation, accumulation, and differences in ex-


cess pore pressure response can reveal the subtleties in mechanical
response between untreated and variously treated calcareous sand
specimens. Fig. 8 shows the development of excess pore pressure with N
for specimens treated with 0.2, 0.4, 0.6 CS and clean sand at a
Fig. 7. CSR versus NL under the effective confining pressure of 50, 100, 200 kPa: (a) CSR=0.25 and effective confining pressure of 100 kPa. The general
Clean sand; (b) Treated (Group TB) sand. response indicated by the pore pressure generation curves are relatively
similar, with a significant increase in the beginning of loading and then
respectively, while the value of b is 0.195 for MICP=0 and 0.180 for a slower rate of accumulation of excess pore pressure until initiation of
MICP=0.2 CS, 0.4 CS, 0.6 CS (as shown in Fig. 6). A single magnitude liquefaction, this phenomena is quite similar to other medium-dense
of b was found to suitability reproduce the CSR trends for all of the calcareous sands [63]. The number of cycles to liquefaction increases
levels of treatment and confining pressures considered in this study, with an increase in the bio-cementation content. The specific number of
which indicates some level of predictability in the effects of the mag- cycles to liquefaction is 31, 72, 118 and 135 for the untreated specimen
nitude of treatment beyond those levels considered herein. and treated specimens with 0.2 CS, 0.4 CS and 0.6 CS, respectively.
Comparison of untreated and treated specimens with 0.4 L CS in-
dicates a five-fold increase in NL is possible at a given CSR (e.g., com- 4.4. Deformation response
pare NL for CSR=0.25 for the untreated specimens to NL for the treated
specimens). The behavior observed in Fig. 6 is attributed to the in- The magnitude of bio-cementation was observed to control the de-
creased strength due to enhancement of cementation during cyclic formation response of specimens. Fig. 9 shows that the deformation
shear of the bio-cemented particles and calcite precipitated in the pore characteristics changed significantly as the magnitude of CS content
space. On the other hand, the increased strength and decreased void increased. For example, the untreated specimen produces compression
ratio resulting from cementation reduced the magnitude of pore pres- and extension strains exceeding 1.5%, with extension strains reaching
sure during cyclic shear, providing additional cyclic resistance [57]. 5% following liquefaction. Specimens prepared with a biocement con-
The effective confining pressure of a given liquefiable soil layer in tent of 0.2 L showed a significantly reduced tendency for compression
part governs its liquefaction resistance [58,59]. Specimens in Group U
and TB, with 0.4 L CS contents, were cyclically tested at σ′c = 50,
100, 200 kPa and at various CSRs to explore the role of effective con-
fining pressure. Fig. 7 demonstrate that the number of cycles required
to initiate liquefaction at a given CSR increases as the effective con-
fining pressure decreases. Similarly, the fitted curves are expressed by
Eq. (3), and the value of a and b are showed in Fig. 7. Considering a
CSR=0.25, the number of cycles required to cause liquefaction in-
creases from 11 to 41 cycles as the effective confining pressure de-
creases from 200 to 50 kPa in the untreated sand. The number of cycles
to liquefaction at the same CSR for the treated specimens increases from
78 to 135 as the effective confining pressures decreases from 200 to
50 kPa. Note, particle size analyses conducted following cyclic shearing
indicated that crushing of particles did not occur for the confining
pressures investigated, similar to findings reported elsewhere [60]. For
treated specimens at CSR=0.33, NL increases from 18 to 49 cycles as
the effective confining pressure decreases from 200 to 50 kPa. The in-
Fig. 9. Axial strain versus N for calcareous sand treated by different CS content under
crease in confining pressure results in a reduction in the cyclic strength
100kPa effective confining pressure.

15
P. Xiao et al. Soil Dynamics and Earthquake Engineering 107 (2018) 9–19

strains, with a maximum compression strain of 1%, and requiring sig-


nificantly greater N to achieve this magnitude of strain. Thereafter, the
compression strain in the specimen reduced, as opposed to continuing
to accumulate and indicating the onset of dilation hardening. As the CS
content continues to increase, the amplitude of cyclic strain decreases
and the tendency for and magnitude of compression decreases. This
behavior indicates that the specimens have gradually transitioned from
a significantly porous material to a more solid material, increasing its
stiffness and limiting its deformation potential. Additionally, this be-
havior is also attributed to the enhancement of the dilative character-
istics produced by MICP, attributed to the introduction of CaCO3 pre-
cipitation in the pore space. Examination of the specimen with 0.6 L CS
following cycling indicated that it remained largely intact despite the
generation of excess pore pressure equal to its confining pressure, and
only small spalls were observed directly adjacent to the loading platens.
Similar observations on cemented specimens have been observed by
Fig. 11. Improvement factor of liquefaction resistance against moment magnitude of
previous researchers [64]. Accordingly, bio-cemented calcareous sands
different bio-cemented samples.
appear significantly improved in their ability to resist long duration,
strong ground motions.
5.2. Correlation between the improvement in cyclic strength and moment
magnitude
5. Discussion
The improvement in liquefaction resistance of bio-cemented sands
5.1. Correlation between NL and N5% as observed in the UCTs can be expressed through an improvement
factor, If, defined as the ratio of the CSR for initial liquefaction of
The number of cycles to liquefaction reported above was identified treated specimens and that of untreated counterpart specimens [70].
for the excess pore pressure reaching a value equal to or near to the Separately, the moment magnitude, Mw, may be related to equivalent
initial effective confining pressure (i.e., ru > 0.95), termed initial li- loading cycles [71]. An earthquake with Mw=6.0 would be expected to
quefaction. Another criterion for failure is set to the number of cycles to generate five significant uniform stress cycles, whereas ten cycles cor-
the development of 5% double amplitude (D.A.) strain, N5% [8,65,66]. respond to Mw=7.0, fifteen cycles represents a Mw=7.5, and twenty-
The relationship between initial liquefaction and 5% D.A. strain is two cycles correspond to Mw=8.0. Based on the correlation between
linear for clean sand, sand-fines mixture and sand treated with a low magnitude and the number of significant cycles, it has become common
cementation [67,68]. Fig. 10 shows the relationships between NL and to relate the magnitude of the improvement in CSR to the number of
N5%, which indicates largely linear correlations, despite the observation significant cycles or Mw [66,72,73]. Fig. 11 compares the relationship
that NL occurred well before 5% DA strain, consistent with previously between If of the biocemented sand and Mw for various levels of
reported findings for uncemented calcareous sands [7,69]. Fig. 10 also treatment at σ′c = 100kPa . The improvement factor increases as the
shows that the slope of the correlation between NL and N5% increases amount of bio-cementation increases; for example, as the amount of
from 1.028 to 1.985 as the CS content increases from 0 to 0.6. It can be treatment increases from 0.2 to 0.6 L, If increases from approximately
interpreted that the soil with greater bio-cementation exhibits a greater 1.10 to about 1.35.
stiffness and strength, resulting in a larger number of cycles for
achieving 5% axial strain. These findings are in line with cyclic tests on 5.3. Physical aspects of induced calcite crystals
other cemented materials [68].
In order to further investigate the impact of CS content on the
crystal formation of MICP-treated samples, SEM analysis was carried
out on specimens from Groups TA, TB, and TC as shown in Fig. 12.
These images were selected from the middle portion of the post-shear
specimens with similar grain sizes for comparative purposes. The
crystals formed mainly on the surface of the sand particles to coat the
particle surfaces, and some portions of the original sand surface can be
observed in the lightly treated specimen (i.e., white circles in
Fig. 12(a)). As shown in Fig. 12(a) through (c), the void space of the
specimen decreases with the increases of bio-cementation in the spe-
cimen. These observations are markedly different from previous studies
on microbial induced calcite precipitation in silica sand [74,75]. Lin
et al. [76] observed that three distinct spatial distributions of CaCO3
precipitate in silica sand during MICP: (a) contact cementing; (b) grain
coating; and (c) matrix supporting. For the calcareous sand in this
study, the spatial distributions of CaCO3 precipitation are mostly dis-
tributed on the sand surface (in grain coating) with increasing matrix
support as the amount of bio-cementation increased (as shown in the
yellow cycles). Similar results were observed by van Paassen [77].
As the magnitude of bio-cementation increases, the particle surface,
inner pore and the gap between particles are gradually filled by the
CaCO3 precipitate. This manifestation lends difficulty in distinguishing
Fig. 10. Comparison between number of cycles to liquefaction determined using initial the boundary between the precipitate and sand particles as shown in
liquefaction criterion versus 5% axial strain double amplitude criterion.
Fig. 12(c). The smooth depressions in the precipitate circled in blue

16
P. Xiao et al. Soil Dynamics and Earthquake Engineering 107 (2018) 9–19

response of clean calcareous sand. Specifically, the pore pressure


response illustrates greater local stability as the bio-cement content
increases, and deformation response exhibited significant reduction
in compressive strains. With regard to resilience under earthquake
loading, the improvement factor of the bio-cemented calcareous
sand reached as large as 1.34 for a moment magnitude 7.5 earth-
quake.
2. The magnitude of cyclic shear and confining pressure plays an im-
portant role in affecting liquefaction resistance. An increase in
confining pressure leads to a decrease in the liquefaction resistance
for both clean and bio-cemented specimens. Similarly, the increase
cyclic shear stress results in a lower magnitude of cycles to lique-
faction. However, the number of cycles to liquefaction increased
with increasing cementation content, and showed that MICP-treat-
ment can serve to significantly reduce liquefaction potential of
calcareous sand.
3. The microscopic images show that the MICP-treatment results in
widespread coating of the calcareous sand particles. As the degree of
cementation increase, the number and size of crystals increases,
explaining the corresponding rate of improvement in cyclic re-
sistance. The improved resistance to liquefaction results from both
the increase in surface roughness and the filling of voids between
particles. For specimens treated with higher amount of cementation
solution, it became difficult to distinguish the boundary between the
precipitation and the sand particles.

The study has shown that microbial induce carbonate precipitation


(MICP) can serve to mitigate liquefaction of calcareous sand. Given the
similarity between MICP treatment and its byproduct (i.e., the bio-ce-
mentation) and the natural deposition of CaCO3 minerals, MICP-treat-
ment can serve as a suitable technology to manufacture artificial coral
reefs and calcareous deposits that can limit the consequences of lique-
faction and control erosion. In addition, a constitutive model based on
Fig. 12. SEM of CaCO3 crystals formed for different MICP-treated samples: (a) Group TA; bounding surface plasticity [82,83] would be introduced in the future
(b) Group TB; (c) Group TC. work to predict cyclic and liquefaction behaviors of the bio-cemented
calcareous sand.
indicate the previous presence of a sand particle in the spatial dis-
tributions of contacting cementing. The widespread CaCO3 precipita- Acknowledgements
tion and its contact spatial distributions can be attributed to the com-
bination of bacterial attachment on sand surfaces [78,79] and the We would like to acknowledge that this study is supported by the
abiotic precipitation processes promoted by the carbonate component National Natural Science Foundation of China (Grant No. 51578096,
of the sand particle [80]. From the magnified SEM images, and focusing Grant No. 51678094 and Grant No. 51509024) and NSF (Grant No.
on the near-field of view, it may be generally observed that the size of CMMI-1538460).
the CaCO3 crystals increased with increases in the CS content; this most
pronounced in Fig. 12(c). Somani et al. [81] found that the average References
crystal size was positively correlated to carbonate concentration.
Hence, the larger size in CaCO3 crystals observed in the 0.6 CS spe- [1] Schwiderski EW. On charting global ocean tides. Rev Geophys 1980;18(18):243–68.
[2] Singh SC, Carton H, Tapponnier P, Hananto ND, Chauhan APS, Hartoyo D, Bayly M,
cimen is a result of the increased carbonate concentration (Fig. 3(b)).
Moeljopranoto S, Bunting T, Christie P. Seismic evidence for broken oceanic crust in
Thus, there appears to be a contribution in surface roughness from the the 2004 Sumatra earthquake epicentral region. Nat Geosci 2008;1(11):777–81.
crystal size to the increase in the liquefaction resistance of treated [3] Salehzadeh H, Procter DC, Merrifield CM. Medium dense non-cemented carbonate
specimens in addition to the improved void reduction and bond re- sand under reversed cyclic loading. Int J Civil Eng 2006.
[4] Wang XZ, Jiao YY, Wang R, Hu MJ, Meng QS, Tan FY. Engineering characteristics of
sistance. the calcareous sand in Nansha Islands, South China Sea. Eng Geol 2011;120:40–7.
[5] Vahdani S, Pyke R, Siriprusanen U. Liquefaction of calcareous sands and lateral
6. Conclusion spreading experienced in Guam as a result of the 1993 Guam earthquake. In: 5th U.
S.- Japan workshop on Earthquake resistant design of lifeline facilities and coun-
termeasures against soil liquefaction., Snowbird, UT, United States; 1994, p.
In this study, a series of undrained cyclic triaxial tests have been 117–123.
conducted to investigate the effect of microbial induce calcite pre- [6] Valle C, Camacho BI, Stokoe KH, Rauch AF. Comparison of the Dynamic Properties
and Undrained Shear Strengths of Offshore Calcareous Sand and Artificially
cipitation on the liquefaction mitigation of calcareous sand. Different Cemented Sand. In: ASME 2003 International Conference on Offshore Mechanics
specimens, including sand treated with 0.2 L, 0.4 L, 0.6 L CS and un- and Arctic Engineering; 2003, p. 133–141.
treated clean sand were tested. Additionally, the effect of different [7] Salem M, Elmamlouk H, Agaiby S. Static and cyclic behavior of North Coast cal-
careous sand in Egypt. Soil Dyn Earthq Eng 2013;55:83–91.
confining pressure and cyclic stress ratio have been compared. The key
[8] Hyodo M, Hyde AFL, Aramaki N. Liquefaction of crushable soils. Geotechnique
findings developed from this research include: 1998;48:527–43.
[9] Coop MR, Sorensen KK, Freitas TB, Georgoutsos G. Particle breakage during
shearing of a carbonate sand. Geotechnique 2004;54(3):157–63.
1. The liquefaction resistance, pore pressure development, and strain
[10] Xiao Y, Liu H, Chen Q, Ma Q, Xiang Y, Zheng Y. Particle breakage and deformation
accumulation of bio-cemented sand deviate significantly from the of carbonate sands with wide range of densities during compression loading

17
P. Xiao et al. Soil Dynamics and Earthquake Engineering 107 (2018) 9–19

process. Acta Geotech 2017. http://dx.doi.org/10.1007/s11440-11017-10580-y. ASTM; 2011.


[11] Xiao Y, Liu H, Xiao P, Xiang J. Fractal crushing of carbonate sands under impact [48] Díaz-Rodríguez JA, Antonio-Izarraras VM, Bandini P, López-Molina JA. Cyclic
loading. Geotech Lett 2016;6(3):199–204. strength of a natural liquefiable sand stabilized with colloidal silica grout. Can
[12] Brown RE. Vibro flotation compaction of cohesionless soils. J Geotech Eng Div Geotech J 2008;45:1345–55.
1977;103(12):1437–51. [49] Achal V, Mukherjee A, Basu PC, Reddy MS. Lactose mother liquor as an alternative
[13] Stuedlein AW, Gianella TN, Canivan G. Densification of granular soils using con- nutrient source for microbial concrete production by Sporosarcina pasteurii. J Ind
ventional and drained timber displacement piles. J Geotech Geoenviron Eng Microbiol Biot 2009;36(3):433–8.
2016;142(12):04016075. [50] ASTM_D5311, Standard test method for load controlled cyclic triaxial strength of
[14] Gianella TN, Stuedlein AW. Performance of driven displacement pile-improved soil. In: American Society of Testing and Materials, West Conshohocken,
ground in controlled blasting field tests. J Geotech Geoenviron Eng Pennsylvania, USA; 2013. p. 1–11.
2017;143(9):04017047. [51] Hammes F, Verstraete W. Key roles of pH and calcium metabolism in microbial
[15] Boulanger RW, Hayden RF. Aspects of compaction grouting of liquefiable soil. J carbonate precipitation. Rev Environ Sci Bio/Technol 2002;1(1):3–7.
Geotech Eng 1995;121(12):844–55. [52] Martinez BC, DeJong JT, Ginn TR, Montoya BM, Barkouki TH, Hunt C, Tanyu B,
[16] Poulos HG, Randolf MF, Semple RM. Evaluation of pile friction from conductor Major D. Experimental optimization of microbial-induced carbonate precipitation
tests. In: the International Conference on Calcareous Sediments, Perth; 1988. p. for soil Improvement. J Geotech Geoenviron Eng 2013;139:587–98.
599–605. [53] Li B., Geotechnical properties of biocement treated sand and clay. In: Civil and
[17] Alba J., Audibert J., Pile design in calcareous and carbonaceous granular materials, Environmental Engineering, Nanyang Technological University, Singapore; 2015.
and historic review. In: Proceedings of the 2nd international conference on en- [54] Choi S-G, Park S-S, Wu S, Chu J. Methods for calcium carbonate content mea-
gineering for calcareous sediments. Rotterdam: AA Balkema; 1999. p. 29–44. surement of biocemented soils. J Mater Civil Eng 2017;29(11):06017015.
[18] Xiao Y, Liu H, Chen Q, Long L, Xiang J. Evolution of particle breakage and volu- [55] Saxena S, Reddy R. Mechanical behavior of cemented sands. Report to the National
metric deformation of binary granular soils under impact load. Granul Matter 2017. Science Foundation Department of Civil Engineering, Illinois Institute of
http://dx.doi.org/10.1007/s10035-10017-10756-z. Technology; 1987. [Report No. IIT-CE-8701].
[19] Wang YH, Leung SC. Characterization of cemented sand by experimental and nu- [56] Price AB, DeJong JT, Boulanger RW. Cyclic loading response of silt with multiple
merical investigations. J Geotech Geoenviron 2008;134(7):992–1004. loading events. J Geotech Geoenviron Eng 2017;143(10):04017080.
[20] Xiao Y, Stuedlein AM, Chen Q, Liu H, Liu P. Stress-strain-strength response and [57] Suazo G, Fourie A, Doherty J. Cyclic shear response of cemented paste backfill. J
ductility of Gravels Improved by polyurethane foam adhesive. J Geotech Geotech Geoenviron Eng 2017;143(1).
Geoenviron 2017:143. http://dx.doi.org/10.1061/(ASCE)GT.1943-5606.0001812. [58] Seed HB. Soil liquefaction and cyclic mobility for level ground during earthquakes.
[21] John W. An introduction to geotechnical processes. Brill; 2005. J Geotech Eng Div 1979;105(2):201–55.
[22] Stocks-Fischer S, Galinat JK, Bang SS. Microbiological precipitation of CaCO3. Soil [59] Dobry R, Ladd R, Yokel F, Chung R, Powell D. Prediction of pore water pressure
Biol Biochem 1999;31:1563–71. buildup and liquefaction of sands during earthquakes by the cyclic strain method,
[23] Whiffin VS, van Paassen La, Harkes MP. Microbial carbonate precipitation as a soil building science series. Washington, DC: National Bureau of Standards, US
improvement technique. Geomicrobiol J 2007;24:417–23. Department of Commerce, US Governmental Printing Office; 1982.
[24] Ivanov V, Chu J. Applications of microorganisms to geotechnical engineering for [60] Petropoulos K., Earthquake induced liquefaction susceptibility of Carbonate sands:
bioclogging and biocementation of soil in situ. Rev Environ Sci Biotechnol Experimental study: Cyclic behaviour of Carbonate materials used as hydraulic fill.
2008;7:139–53. In: Civil engineering and Geosciences, Delft University of Technology; 2013.
[25] DeJong JT, Mortensen BM, Martinez BC, Nelson DC. Bio-mediated soil improve- [61] Suazo G, Fourie A, Doherty J, Hasan A. Effects of confining stress, density and initial
ment. Ecol Eng 2010;36:197–210. static shear stress on the cyclic shear response of fine-grained unclassified tailings.
[26] Krajewska B. Ureases I. Functional, catalytic and kinetic properties: a review. J Mol Geotechnique 2016;66:401–12.
Catal B: Enzym 2009;59(1–3):9–21. [62] Keramatikerman M, Chegenizadeh A, Nikraz H. Experimental study on effect of fly
[27] Burne RA, Chen YYM. Bacterial ureases in infectious diseases. Microbes Infect ash on liquefaction resistance of sand. Soil Dyn Earthq Eng 2017;93:1–6.
2000;2(5):533–42. [63] Ghionna VN, Porcino D. Liquefaction resistance of undisturbed and reconstituted
[28] Nemati M, Voordouw G. Modification of porous media permeability, using calcium samples of a natural coarse sand from undrained cyclic triaxial tests. J Geotech
carbonate produced enzymatically in situ. Enzym Microb Technol Geoenviron Eng 2006;132(2):194–202.
2003;33(5):635–42. [64] Porcino D, Marcianò V, Granata R. Undrained cyclic response of a silicate-grouted
[29] Chu J, Stabnikov V, Ivanov V, Li B. Microbial method for construction of an sand for liquefaction mitigation purposes. Geomech Geoengin 2011;6:155–70.
aquaculture pond in sand. Geotechnique 2013:1–5. [65] Gallagher PM, Mitchell JK. Influence of colloidal silica grout on liquefaction po-
[30] Tittelboom KV, Belie ND, Muynck WD, Verstraete W. Use of bacteria to repair tential and cyclic undrained behavior of loose sand. Soil Dyn Earthq Eng
cracks in concrete. Cem Concr Res 2010;40(1):157–66. 2002;22:1017–26.
[31] Achal V, Mukerjee A, Reddy MS. Biogenic treatment improves the durability and [66] Park S-S, Kim Y-S. Liquefaction resistance of sands containing plastic fines with
remediates the cracks of concrete structures. Constr Build Mater 2013;48(19):1–5. different plasticity. J Geotech Geoenviron Eng 2013;139:825–30.
[32] DeJong JT, Fritzges MB, Nüsslein K. Microbially induced cementation to control [67] Ochoa-Cornejo F, Bobet A, Johnston CT, Santagata M, Sinfield JV. Cyclic behavior
sand response to undrained shear. J Geotech Geoenviron Eng 2006;132:1381–92. and pore pressure generation in sands with laponite, a super-plastic nanoparticle.
[33] van Paassen La, Daza CM, Staal M, Sorokin DY, van der Zon W, van Loosdrecht Soil Dyn Earthq Eng 2016;88:265–79.
MCM. Potential soil reinforcement by biological denitrification. Ecol Eng [68] Maker MH, Ro KS, Welsh JP. Cyclic undrained behavior and liquefaction potential
2010;36:168–75. of sand treated with chemical grouts and microfine cement (MC-500). Geotech Test
[34] Martinez BC, Dejong JT. Bio-mediated soil improvement: load transfer mechanisms J 1994;17:159–70.
at the micro- and macro- scales. Geotech Spec Publ 2009;188:242–51. [69] LaVielle TH. Liquefaction susceptibility of uncemented calcareous sands from
[35] Montoya BM, DeJong JT. Stress-strain behavior of sands cemented by microbially Puerto Rico by cyclic triaxial testing. Virginia Polytechnic Institute and State
induced calcite precipitation. J Geotech Geoenviron Eng 2015;141(6):04015019. University; 2008. p. 177.
[36] Feng K, Montoya BM. Influence of confinement and cementation level on the be- [70] Porcino D, Marcianò V, Granata R. Cyclic liquefaction behaviour of a moderately
havior of microbial-induced calcite precipitated sands under monotonic drained cemented grouted sand under repeated loading. Soil Dyn Earthq Eng
loading. J Geotech Geoenviron Eng 2016;142(1):04015057. 2015;79:36–46.
[37] Han Z, Cheng X, Qiang M. An experimental study on dynamic response for MICP [71] Idriss I, Boulanger RW. Soil liquefaction during earthquakes. San Francisco:
strengthening liquefiable sands. Earthq Eng Eng Vib 2016;15(4):673–9. Earthquake Engineering Research Institute; 2008.
[38] Burbank M, Weaver T, Lewis R, Williams T, Williams B, Crawford R. Geotechnical [72] El Takch A, Sadrekarimi A, El Naggar H. Cyclic resistance and liquefaction behavior
tests of sands following bioinduced calcite precipitation catalyzed by indigenous of silt and sandy silt soils. Soil Dyn Earthq Eng 2016;83:98–109.
bacteria. J Geotech Geoenviron Eng 2013;139:928–36. [73] James M, Aubertin M, Wijewickreme D, Wilson GW. A laboratory investigation of
[39] Feng K, Montoya BM. Quantifying level of microbial-induced cementation for cy- the dynamic properties of tailings. Can Geotech J 2011;48:1587–600.
clically loaded sand. J Geotech Geoenviron Eng 2017:06017005. [74] Chou C-W, Seagren Ea, Aydilek AH, Lai M. Biocalcification of Sand through
[40] Montoya B, DeJong J, Boulanger R. Dynamic response of liquefiable sand improved Ureolysis. J Geotech Geoenviron Eng 2011;137:1179–89.
by microbial-induced calcite precipitation. Geotechnique 2013:302–12. [75] Cheng L, Shahin MA, Cord-Ruwisch R. Bio-cementation of sandy soil using micro-
[41] Sasaki T, Kuwano R. Undrained cyclic triaxial testing on sand with non-plastic fines bially induced carbonate precipitation for marine environments. Geotechnique
content cemented with microbially induced CaCO3. Soils Found 2016;56:485–95. 2014;64(12):1010–3.
[42] Sharma SS, Ismail Ma. Monotonic and cyclic behavior of two calcareous soils of [76] Lin H, Suleiman MT, Brown DG, Kavazanjian E. Mechanical behavior of sands
different origins. J Geotech Geoenviron Eng 2006;132:1581–91. treated by microbially induced carbonate precipitation. J Geotech Geoenviron Eng
[43] Airey DW, Fahey M. Cyclic response of calcareous soil from the North-West Shelf of 2016;142(2):04015066.
Australia. Geotechnique 1991;41:101–21. [77] Paassen L.A.V., Biogrout, ground improvement by microbial induced carbonate
[44] Khan MNH, Shimazaki S, Kawasaki S. Coral sand solidification test through mi- precipitation. In: Delft University of Technology The Nethelands; 2009.
crobial calcium carbonate precipitation using Pararhodobacter sp. Int J Geomate [78] Scholl MA, Mills AL, Herman JS, Hornberger GM. The influence of mineralogy and
2016. solution chemistry on the attachment of bacteria to representative aquifer mate-
[45] Khan MNH, Amarakoon GGNN, Shimazaki S, Kawasaki S. Coral sand solidification rials. J Contam Hydrol 1990;6(4):321–36.
test based on microbially induced carbonate precipitation using ureolytic bacteria. [79] Shellenberger K, Logan BE. Effect of molecular scale roughness of glass beads on
Mater Trans 2015;56:1725–32. colloidal and bacterial deposition. Environ Sci Technol 2002;36(2):184–9.
[46] Hassanlourad M, Salehzadeh H, Shahnazari H. Mechanical properties of ungrouted [80] Lioliou MG, Paraskeva CA, Koutsoukos PG, Payatakes AC. Heterogeneous nuclea-
and grouted carbonate sands. Int J Geotech Eng 2010;4:507–16. tion and growth of calcium carbonate on calcite and quartz. J Colloid Interface Sci
[47] ASTM_D2487. Standard practice for classification of soils for engineering purposes. 2007;308(2):421–8.

18
P. Xiao et al. Soil Dynamics and Earthquake Engineering 107 (2018) 9–19

[81] Somani RS, Patel KS, Mehta AR, Jasra RV. Examination of the polymorphs and soils. Int J Geomech 2017;17(7):04016150.
particle size of calcium carbonate precipitated using still effluent (i.e., CaCl2 + [83] Xiao Y, Sun Y, Liu H, Xiang J, Ma Q, Long L. Model predictions for behaviors of
NaCl Solution) of soda ash manufacturing process. Ind Eng Chem Res sand-nonplastic-fines mixturesusing equivalent-skeleton void-ratio state index. Sci
2006;45(15):5223–30. China Technol Sci 2017;60(6):878–92.
[82] Xiao Y, Sun Y, Yin F, Liu H, Xiang J. Constitutive modeling for transparent granular

19

View publication stats

You might also like