You are on page 1of 13

Mechanical Behavior of Sands Treated by Microbially

Induced Carbonate Precipitation


Hai Lin, S.M.ASCE 1; Muhannad T. Suleiman, A.M.ASCE 2; Derick G. Brown, M.ASCE 3;
and Edward Kavazanjian Jr., F.ASCE 4
Downloaded from ascelibrary.org by Technische Universiteit Delft on 03/28/24. Copyright ASCE. For personal use only; all rights reserved.

Abstract: The mechanical behavior of sands treated using microbially induced carbonate precipitation (MICP) has been investigated at the
macroscale and the microscale. Triaxial and confined compression tests with embedded shear and compression wave (S-wave and P-wave)
sensors were conducted on two MICP-treated sands, Ottawa 50=70 and 20=30 silica sands. Triaxial compression tests were conducted at three
different confining pressures (25, 50, and 100 kPa). Tests were also performed at calcium chloride (CaCl2 ) concentrations of 0.1 and 0.3 M,
resulting in specimens with average calcium carbonate (CaCO3 ) content ranging from 1.5 to 2.5% for the 50=70 sand and from 1 to 1.6% for
the 20=30 sand. In contrast to previous research, the results of triaxial tests presented in this paper show an increase of the soil strength even at
1% calcium carbonate content. After the tests, samples taken from the specimens were utilized to measure the CaCO3 content and to perform
analysis using scanning electron microscopy (SEM) and energy dispersive X-ray spectroscopy (EDS). The SEM and EDS images were used
to assess the morphology and spatial distributions of CaCO3 at the microscale. DOI: 10.1061/(ASCE)GT.1943-5606.0001383. © 2015
American Society of Civil Engineers.
Author keywords: Bio-mediated soil improvement; Calcium carbonate; Cementation; Microbially induced carbonate precipitation.

Background the equilibrium of calcium carbonate (CaCO3 ) precipitation/


dissolution toward precipitation by increasing the availability
Biomediated soil improvement is an innovative ground improve- of the carbonate ion (Stocks-Fischer et al. 1999; Ebigbo et al.
ment solution that could be suitable for many geotechnical prob- 2012).
lems (Mitchell and Santamarina 2005; Ivanov and Chu 2008; The precipitated calcium carbonate can coat soil particles, ce-
Kavazanjian and Karatas 2008; DeJong et al. 2010a, 2013). One ment the soil matrix, and fill the soil void space, increasing its
promising biomediated technique that has attracted a lot of atten- strength, stiffness, and dilatancy. MICP has been investigated
tion recently is microbially induced carbonate precipitation (MICP) for several geotechnical applications including soil improvement,
(Dejong et al. 2006; Whiffin et al. 2007; Karatas et al. 2008; van liquefaction mitigation, stabilizing costal sand dunes, radionu-
Paassen 2009; Hamdan et al. 2011; Al Qabany et al. 2012; Burbank clide sequestration, and fugitive dust stabilization (Whiffin et al.
et al. 2013). 2007; Bang et al. 2009; Kavazanjian et al. 2009; van der Ruyt
The most common MICP process described in the technical and van der Zon 2009; Fujita et al. 2010; Burbank et al. 2013;
literature is hydrolysis of urea by the enzyme urease (ureolysis). Montoya et al. 2013). In spite of the considerable interest in
The microorganism most commonly employed for MICP through MICP, available information on the mechanical behavior of
ureolysis is Sporosarcina pasteurrii (ATCC 11859 and DSM-33). MICP-treated sands on both the macroscale and the microscale
Sporosarcina pasteurrii is an alkalophilic soil bacterium with a is still limited.
highly active urease enzyme (Ferris et al. 1996). The bacteria Recently, Lin et al. (2015) investigated the use of MICP to
produce urease to hydrolyze urea into ammonium and carbonic improve the static axial capacity of permeable piles focusing on
acid, which is accompanied by an increase of alkalinity (the pervious concrete piles. Lin et al. (2015) reported that the CaCO3
logarithmic of acid dissociation constant, pKa of ammonia/ content of the soil surrounding the tested pervious concrete
ammonium is 9.3 and pH of ∼9). This alkaline environment shifts piles ranged from 0 to 3%. Therefore, the triaxial tests reported
in this paper focus on the drained response of MICP-mediated
1
Graduate Research Assistant, Dept. of Civil and Environmental sands with CaCO3 contents in the range reported by Lin et al.
Engineering, Lehigh Univ., 390 STEPs Bldg., 1 W Packer Ave., (2015).
Bethlehem, PA 18015.
2
Associate Professor, Dept. of Civil and Environmental Engineering,
Macroscale Behavior of MICP-Treated Sand
Lehigh Univ., 326 STEPs Bldg., 1 W Packer Ave., Bethlehem, PA
18015 (corresponding author). E-mail: mts210@lehigh.edu At the macroscale, the mechanical behavior of MICP-treated sands
3
Associate Professor, Dept. of Civil and Environmental Engineering, is controlled by the CaCO3 content, sand gradation, relative den-
Lehigh Univ., 346 STEPs Bldg., 1 W Packer Ave., Bethlehem, PA 18015. sity, and confining stress (Rebata-Landa 2007; Chou et al. 2011;
4
Professor, Dept. of Civil Engineering, School of Sustainable van Paassen et al. 2012; Al Qabany and Soga 2013; Cheng et al.
Engineering and the Built Environment, Arizona State Univ., Tempe,
2013). However, most researchers have been focusing on investigat-
AZ 85287-5306.
Note. This manuscript was submitted on July 18, 2014; approved on June ing the behavior of one type of MICP-treated sand under undrained
5, 2015; published online on July 28, 2015. Discussion period open until condition, providing limited information on the variations of volume
December 28, 2015; separate discussions must be submitted for individual change, modulus, friction angle, and cohesion. In addition, very
papers. This paper is part of the Journal of Geotechnical and Geoenviron- limited drained triaxial stress-strain responses have been reported
mental Engineering, © ASCE, ISSN 1090-0241/04015066(13)/$25.00. in the literature for MICP-treated specimens at different CaCO3

© ASCE 04015066-1 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2016, 142(2): 04015066


into the pore space and eventually create a CaCO3 bridge between
sand particles (matrix supporting). Fig. 1 shows a schematic illus-
tration of the contact cementing, grain coating, and matrix support-
ing mechanisms. Contact cementing and grain coating were
observed in scanning electron microscopy (SEM) and/or energy
dispersive X-ray spectroscopy (EDS) images by several researchers
(Martinez and DeJong 2009; van Paassen 2009; Dejong et al.
(a) (b) (c)
2010b; Al Qabany et al. 2012). However, observation of the matrix
Fig. 1. Ideal spatial distributions of CaCO3 during MICP: (a) contact supporting mechanism has not been reported in the literature for
cementing; (b) grain coating; (c) matrix supporting (note: figures not to MICP-treated sands.
scale) The goal of this paper is to investigate the macroscale mechani-
cal behavior of MICP-treated sands under drained triaxial compres-
Downloaded from ascelibrary.org by Technische Universiteit Delft on 03/28/24. Copyright ASCE. For personal use only; all rights reserved.

sion and under zero lateral strain (i.e., 1D confined compression)


conditions and to investigate the spatial distribution of CaCO3 at
contents. Furthermore, tests performed on MICP-treated sands
the microscale. Triaxial and confined compression tests with em-
under confined compression condition (one-dimensional [1D]
bedded shear and compression wave (S-wave and P-wave) sensors
compression loading) are limited in the literature. Therefore, the
were conducted on two MICP-treated sands, Ottawa 50/70 and 20/
focus of the tests reported in this paper is to investigate the evolution
of strength, stiffness, volumetric strain as a function of confinement, 30 silica sands. The triaxial compression tests were conducted at
axial strain, and CaCO3 cementation level for different sand three different confining pressures (25, 50, and 100 kPa). Tests
types under consolidated drained (CD) and 1D (Ko ) loading were also performed at different CaCO3 cementation contents tar-
conditions. geting the range reported by Lin et al. (2015) surrounding pervious
concrete piles. After the tests, samples at different locations within
the specimens were saved for CaCO3 content measurement and
Microscale of MICP-Treated Sand SEM and EDS imaging. The morphology and spatial distribution
At the microscale, the precipitated calcium carbonate has been of CaCO3 at the microscale were analyzed using the SEM and EDS
observed to nucleate around bacteria cells attached to the sand results.
particle surface, forming calcite and vaterite CaCO3 crystals or
amorphous CaCO3 (Burbank et al. 2013). The types of CaCO3
isomorphs formed in the precipitation process are primarily deter- Equipment
mined by the urea hydrolysis rate (van Paassen 2009). For example,
spherical vaterite and amorphous CaCO3 are dominant under Triaxial and confined compression (i.e., 1D compression) tests
higher urea hydrolysis rates, whereas rhomboidal calcite formation were used to investigate the mechanical behavior of the two
is dominant under low hydrolysis rates (van Paassen 2009). MICP-treated sands. Fig. 2 presents schematic diagrams of the test
CaCO3 may deposit at or near particle contact points (contact equipment. For the triaxial cell, S-and P-wave sensors were sealed
cementing), coat the soil particles (grain coating), fill the void space inside two fabricated acrylic plates fixed to the top and bottom caps
without contacting a soil particle, or grow from the particle surface [Fig. 2(a)]. The confined compression test device was constructed

Fig. 2. Test setup and instrumentation: (a) triaxial cell; (b) top view of the confined compression cell; (c) side view of confined compression cell
(all dimensions in mm)

© ASCE 04015066-2 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2016, 142(2): 04015066


of a stainless steel cell with two embedded pairs of S-wave sensors at 10% finer by mass (D10 ) of 0.26 mm, a particle diameter at 50%
to measure the S-wave velocity in the vertical and horizontal direc- finer by mass (D50 ) of 0.33 mm, a coefficient of uniformity (Cu ) of
tions [Figs. 2(b and c)]. The S- and P-wave sensors were fabricated 1.43, a coefficient of curvature (Cc ) of 1.01, and maximum and
in-house using piezo elements (4 × 8 mm parallel type-PSI-5H4E minimum void ratio (emax and emin ) of 0.87 and 0.55, respectively.
T226-H4-303Y from Piezo Systems) and unimorph bender discs Ottawa 20=30 sand has a D10 of 0.58 mm, a D50 of 0.71 mm, a Cu
(8.9 mm diameter, 0.11 mm thickness FT-8.9T-9.0A1 from APC of 1.17, a Cc of 1.02, and emax and emin of 0.74 and 0.51, respec-
International, Mackeyville, Pennsylvania). A 2 V square wave with tively. Both sands have greater than 98.7% silica (SiO2 ). To keep
100 HZ frequency provided by a function generator (Agilent the sand surface clean and deplete any soluble chemicals that may
33220A) was used as an input signal for the S-wave sensors. interfere with the CaCO3 content measurement, the sands were ini-
The received signal was filtered by a band-pass filter (Krohn-Hite tially soaked in 1M nitric acid solution for 24 h followed by clean-
3944) of 100 Hz–30 kHz, and a digital oscilloscope (Agilent ing with deionized water and drying in an oven at 105°C for 24 h
DSO5014A) was used to stack 128 signals to reduce noncoherent before being used.
noise (Yun and Santamarina 2005). The S-wave velocity (V s ) was
Downloaded from ascelibrary.org by Technische Universiteit Delft on 03/28/24. Copyright ASCE. For personal use only; all rights reserved.

determined by dividing the tip to tip distance (L in Fig. 2) between


two sensors by the travel time (t) using the first arrival point of the Experimental Procedures
signal (Lee and Santamarina 2005). To generate a P-wave through
the specimen, a 3 V square wave with a frequency of 50 HZ from Sample Preparation
the function generator was used as the input signal. The received
signal is filtered by a band-pass filter of 600 Hz to 300 kHz. The Triaxial test specimens 72 mm in diameter and 145 mm in height
P-wave velocity (V p ) was calculated in a similar manner to the were prepared using the wet-raining method (Chaney and Mulilis
S-wave velocity measurement described above. 1978). Then, a vacuum pressure similar to the target confining pres-
sure during the test (e.g., 25, 50, and 100 kPa) was applied to the
specimens. The average measured void ratio (e) after vacuum con-
Materials solidation was 0.74 and 0.65 for the 50=70 and 20=30 sand spec-
imens, respectively, corresponding to relative densities of 41%
and 39%, respectively. After filling the triaxial cell with water,
Bacteria Cultivation and MICP Recipe the vacuum pressure was decreased by 10 kPa and the cell pressure
The gram-positive bacteria strain Sporosarcina pasteurii (ATCC increased simultaneously by 10 kPa until the cell pressure increased
11859) was used in this study. Bacterial cultures were grown in to the target confining pressure (25, 50, and 100 kPa) and the vac-
an ATCC specified medium (Table 1), harvested in the late expo- uum pressure returned to 0 kPa. In the confined compression tests,
nential growth phase, and then stored in 15% glycerol at −86°C specimens 70 mm in diameter and 60 mm in height were also pre-
using the glass bead method to maintain a uniform bacterial stock pared by wet-raining into the cell to approximately the same rel-
(Jones et al. 1991). Before each experiment, bacteria from the fro- ative densities as the triaxial samples (approximately 40%). Then, a
zen stock were grown in the growth medium inside an incubator 10 kPa normal (compression) pressure was applied on the top cap.
shaker at 170 rpm and 33°C. The bacteria were harvested at It is important to note that all tests were successfully duplicated to
OD600 of 0.8–1.2 (OD600 : optical density of a sample measured verify repeatability and validate the results.
at a wavelength of 600 nm), centrifuged twice at 4,000 g for
30 min to a target bacteria density of 1×108 cells=mL, and stored MICP Treatment
at 4°C until used. Tris buffer and growth medium were used to grow
the bacteria cells. Table 1 presents the chemical recipe that was A peristaltic pump (Cole-Parmer, C/L tubing pump) was used to
used for growing the bacteria cells. introduce chemical and bacteria solutions from the bottom port
of the triaxial and confined compression cells. The flow rate
was set at 10 mL=min. Each specimen was initially flushed by
Sand Types two pore volumes of deionized water (500 mL for triaxial test,
200 mL for confined compression test). Stored bacteria were sus-
Two types of silica sands, Ottawa 50=70 and Ottawa 20=30, were
pended in two pore volumes of urea media prepared in accordance
used in the experiments. Ottawa 50=70 sand has a particle diameter
with the proportions provided in Table 1, stirred, and introduced
into the triaxial and confined compression specimens from the bot-
Table 1. Summary of Microbially Induced Carbonate Precipitation tom port. The bacteria-urea solution was kept within the specimen
(MICP) Recipe for 7 h to allow the bacteria to attach to the sand surface and for
Solution Constituents hydrolysis to occur. After 7 h, two pore volumes of cementation
medium prepared in accordance with the proportions provided
Tris buffer 7.6 g Tris hydrochloric acid
in Table 1 was introduced through the bottom port five times at
54.7 g Tris base in 500 mL distilled water
Growth medium 10 g Yeast extract
3-h intervals. During treatment, the S- and P-wave velocities
5 g Ammonium sulphate in 500 mL of the specimens were monitored with a sampling rate of 1
of 0.13 M Tris buffer (pH 9.0) sample/30 min. After 3 h of finishing the fifth cementation medium,
Sterilized by filter five pore volumes of deionized water were flushed through each
Urea medium 20 g=L Urea specimen from the bottom port to remove residual chemicals.
2.12 g=L NaHCO3
20 g=L NH4 CL
3 g=L Bacto nutrient broth Saturation and Loading
Adjust pH of the medium to 6.0 After MICP treatment, consolidated drained triaxial tests were per-
with 5 M HCL sterile filtration formed. Triaxial specimens were back pressure saturated until the B
Cementation medium Urea medium constituent value (pore water pressure ratio) exceeded 0.95. The S-wave veloc-
0.1 M CaCl2 or 0.3 M CaCl2
ity was monitored during saturation and the changes were less than

© ASCE 04015066-3 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2016, 142(2): 04015066


50 m=s for all tests, which indicates that the saturation process did Table 2. Characteristics of Triaxial Test Specimens
not significantly change either the S-wave velocity or other Effective CaCl2 in
mechanical properties of the sands. Following saturation, the confining cementation Measured
specimens were loaded by displacement control with a loading Untreated stress, σ30 , mediuma, CaCO3
rate of 0.5 mm=min (0.34% axial strain/min) until the axial strain Sand types or MICP (kPa) (mM) contentb, (%)
was approximately equal to 10%. During loading, the S-wave Ottawa 50=70 Untreated 25 0 0
velocity was measured with a sampling rate of 1 sample/0.1% axial Untreated 50 0 0
strain up to an axial strain of 1% and 1 sample/0.35% axial strain Untreated 100 0 0
afterwards. MICP 25 100 1.6
In the confined compression tests, after MICP treatment under MICP 50 100 1.6
10 kPa vertical stress, the vertical (normal) stress was increased to MICP 100 100 1.5
50, 100, 200, 400, 600, 800, 1,000, 1,200, 1,400, and 1,600 kPa MICP 25 300 2.5
and then decreased targeting the same values. Each loading and Ottawa 20/30 Untreated 25 0 0
Downloaded from ascelibrary.org by Technische Universiteit Delft on 03/28/24. Copyright ASCE. For personal use only; all rights reserved.

Untreated 50 0 0
unloading increment lasted for 15 min as recommended by Yun
Untreated 100 0 0
and Santamarina (2005). S-wave velocity and displacement mea- MICP 25 100 1.1
surements were recorded at the end of each loading stage. MICP 50 100 1.0
MICP 100 100 0.9
SEM and EDS Imaging a
From Table 1.
b
Calculated from Eqs. (1) and (2).
After loading, samples were saved for SEM imaging and EDS
analysis. The EDS system was integral to the SEM device (FEI
XL30). The SEM images of the samples were produced first. Then,
EDS was used to scan the sample for Si and Ca elements mapping. Table 3. Characteristics of Confined Compression Test Specimens
Initial effective CaCl2 in Measured
CaCO3 Content Measurement Untreated confining stress, cementation CaCO3
Sand types or MICP σ00 , (kPa) mediuma, (mM) contentb, (%)
After each triaxial and confined compression test, samples of ap-
proximately 15 g in mass were collected at a frequency of one sam- Ottawa 50=70 Untreated 10 0 0
ple every 10 mm of specimen height for CaCO3 content analysis. MICP 10 100 1.4
The samples were placed in glass tubes and oven-dried at 105°C for MICP 10 300 2.6
Ottawa 20=30 Untreated 10 0 0
at least 48 h. After oven drying, 15 mL of 5 M hydrochloric acid
MICP 10 100 0.6
was added to the glass tube. The tubes were capped and shaken MICP 10 300 1.6
gently to facilitate dissolution of the CaCO3 in the sample. The a
liquid samples were extracted and diluted by 1,000 times. The From Table 1.
b
Ca2þ concentration (CCa , g=mL) of the diluted samples were mea- Calculated using Eqs. (1) and (2).
sured using an atomic absorption spectrometer (AAnalyst 200,
PerkinElmer). Then, the CaCO3 content was calculated using
Eqs. (1) and (2) for both sands show an increase following each injection of
the cementation medium. By the end of the MICP treatments, the
W CaCO3 ¼ Cca × 1,000=40 g=mol × 100 g=mol × 15 mL ð1Þ S-wave velocities increased by an average of 1.8 times for the
50=70 sand and 2.6 times for the 20=30 sand. After the end of treat-
W CaCO3 ment, the P-wave velocities increased by an average of 8.7% for the
CaCO3 content ð%Þ ¼ ×100% ð2Þ
W tubeþsand −W tube −W CaCO3 50=70 sand and 10.5% for the 20=30 sand. The smaller increase of
P-wave velocity compared with the S-wave velocity may be attrib-
where W CaCO3 = total weight of CaCO3 in the specimen; W tubeþsand = uted to the dominant influence of the pore water on the bulk modu-
weight of treated sand and glass tube; W tube = weight of the glass lus of saturated specimens. The variation of the S-wave velocity
tube, the factor of 1,000 in Eq. (1) is to account for the 1,000 times during treatment shown in Fig. 3 agrees well with the trend shown
dilution, 40 g=mol is the calcium atomic weight; 100 g=mol = in Dejong et al. (2010b).
CaCO3 molecular weight; and 15 mL = volume of hydrochloric acid
solution added into the glass tube. The detailed information on Consolidated Drained Sand Behavior
sand and treatment types, effective confining pressures, CaCl2 con- The deviator stress versus axial strain for untreated specimens
centrations in cementation media, and average CaCO3 contents under 25, 50, and 100 kPa effective confining pressures reveal
[CaCO3 ðgÞ=sand ðgÞ] after the tests are summarized in Tables 2 strain hardening behavior for both sands [Figs. 4(a and b)]. How-
and 3 for the triaxial and confined compression tests, respectively. ever, the deviator stress versus axial strain curves of the 0.1 M
CaCl2 =MICP-treated specimen show strain softening behavior,
with a post-peak strength decrease. A comparison of untreated
Triaxial Tests and 0.1 M CaCl2 =MICP-treated specimen indicates that the peak
strength of MICP-treated specimen increased by an average of 93%
Sands Treated with 0.1 M CaCl2 for the 50=70 sand and 171% for the 20=30 sand compared with the
peak strength of the untreated specimen. The ultimate strength
S- and P-Wave Velocities during MICP Treatment (defined as the deviator stress at strains ≥ 10%) for both untreated
Figs. 3(a and b) show the variation of S- and P-wave velocities ver- and 0.1 M CaCl2 =MICP-treated specimens are approximately the
sus time during MICP treatment for the 50=70 and 20=30 sands same for the 50=70 sand. However, the ultimate strength of 0.1 M
treated with 0.1 MCaCl2 cementation medium. S-wave velocities CaCl2 =MICP-treated 20=30 sand for the three confining pressures

© ASCE 04015066-4 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2016, 142(2): 04015066


900 2000
ultimate state. A similar residual cohesion was reported by Clough
P wave 1800 et al. (1981) for sands treated by portland cement as discussed
800
1600 before.
700 The initial tangent Young’s moduli (Ei ) were calculated from
S-wave velocity (m/s)

P-wave velocity (m/s)


Confining Pressure (kPa) 1400
600 5th stress-strain curves. These Ei values are presented in Fig. 5(a).
4th
3rd
1200
Ei as a function of confining pressure was evaluated using the
500 1st 1000 power function suggested by Janbu (1963) Eq. (3), where Pa =
2nd
800 atmospheric pressure; k = modulus number; and n = modulus
400
S wave
exponent
600
300 50/70 sand
400 Ei ¼ kPa ðσ3 =Pa Þn ð3Þ
S 25 P 25
200 S 50 P 50
Note: arrows show the onset 200
of cementation medium S 100 P 100 Fig. 5(a) shows that the modulus exponent (n) of the untreated
Downloaded from ascelibrary.org by Technische Universiteit Delft on 03/28/24. Copyright ASCE. For personal use only; all rights reserved.

100 0 sand was 0.53 for the 20=30 sand and 0.88 for the 50=70 sand. For
(a)
0 2 4 6 8 10 12 14 16 18
the 0.1 M CaCl2 =MICP-treated sands, n decreased to 0.09 for the
Time (hrs)
20=30 sand and to 0.19 for the 50=70 sand. These values of n
900 2000 indicate that Ei of MICP-treated sands is less sensitive to the
P wave 1800 change of confining pressure (within the tested range) than un-
800
4th
5th treated sands and may be largely controlled by the level of
1600
700 CaCO3 cementation. The results of initial modulus trend as a func-
S-wave velocity (m/s)

P-wave velocity (m/s)


3rd 1400 tion of effective confining pressure presented in Fig. 5(a) are con-
600 sistent with information provided by Clough et al. (1981) on
1st 1200
2nd
500 1000
naturally-cemented soils and sands cemented by portland cement
and those reported by Yun et al. (2007) on hydrate-bearing
800
400 sediments.
S wave
20/30 sand 600
300
400
S-Wave Velocity during Shear Loading
S 25 P 25
200 Note: arrows show the onset S 50 P 50 Figs. 5(b and c) present the S-wave velocities of the untreated and
of cementation medium S 100 P 100
200
0.1 M CaCl2 =MICP-treated 50=70 and 20=30 sands during triaxial
100 0 compression loading. The S-wave velocities at axial strains of 0 and
0 2 4 6 8 10 12 14 16 18 0.07% were connected with straight lines in these figures based on
(b) Time (hrs)
the assumption of constant small-strain stiffness (Clayton 2011).
Fig. 3. S- and P-wave velocities versus time during MICP treatment: The S-wave velocities were essentially constant for the untreated
(a) for 50=70; (b) 20=30 sands (numbers in the legend are effective specimens. A reduction of S-wave velocities in the 0.1 M
confining pressures in kPa) CaCl2 =MICP-treated specimens was observed starting at an axial
strain between 0.1 and 0.2% in both sands. At an axial strain of
approximately 10%, the S-wave velocities decreased by an average
of 170 and 265 m=s for 50=70 and 20=30 sands, respectively,
ranges from 14 to 41% higher than that of the untreated specimens. which is an average decrease by 41 and 43%, respectively. The
This increase in ultimate strength could be attributed to residual small figure included in Fig. 5(b) shows the change of S-wave
cohesion in the MICP-treated 20/30 sand, as discussed later in this velocities for untreated 50=70 sand at a confining pressure of
25 kPa (a change of less than 10 m=s). Whereas the S-wave veloc-
section, similar to the observation reported by Clough et al. (1981)
ities of the 0.1 M CaCl2 =MICP-treated specimens decreased to-
for sands modified by portland cement.
wards the values for the untreated specimens, the S-wave
Figs. 4(c and d) present the volumetric strain versus axial strain
velocities in the treated specimens at an axial strain of 10% were
behavior for 50=70 and 20=30 sands. MICP treatment with 0.1 M
still higher than that of untreated samples. This could be attributed
CaCl2 significantly increases the dilatancy of both sands. The 0.1
to CaCO3 cementation away from the localized shear-band failure
M CaCl2 =MICP-treated specimen experienced less contraction at
zones observed in the treated specimens. This response is similar to
small strains (in some cases almost no contraction is observed)
the S-wave degradation behavior reported by Tatsuoka and Shibuya
followed by more dilation at large strains than the untreated spec- (1991) on cement-treated sands.
imens. Observation of the specimens failure mode revealed bulging
failures for the untreated specimens and localized shear-band fail- CaCO3 Content of Triaxial Specimens
ures for the 0.1 M CaCl2 =MICP-treated specimens. Fig. 6 presents the measured CaCO3 content along the height of
The Mohr-Coulomb friction angle and cohesion were calculated specimens treated with 0.1 M CaCl2 for both sands. The average
from the triaxial tests results. The peak (equal to ultimate) friction CaCO3 contents of the 50=70 and 20=30 sands were 1.6 and 1%,
angles are 35 and 32° for untreated 50=70 and 20=30 sands, respec- respectively (Table 2). The finer grained 50=70 sand under 25 and
tively. For samples treated with 0.1 M CaCl2 , the peak friction 50 kPa confinement shows a gradient of CaCO3 content along the
angles are 32° for the 50=70 sand and 31° for the 20=30 sand with height of the specimens. However, the 50=70 sand under 100 kPa
associated cohesions of 41 and 58 kPa for 50=70 and 20=30 sands, confinement shows a parabolic shape with higher CaCO3 content
respectively. In addition, the ultimate friction angles are 36° for the at the middle section. It is noted that the void ratios of 50=70 sand
50=70 sand and 33° for the 20=30 sand. The cohesions at the ulti- specimens are similar under 25, 50, and 100 kPa confinement. The
mate strength were 1 and 7 kPa for MICP-treated 50=70 and 20=30 two types of CaCO3 content profiles (linear and parabolic) are sim-
sands, respectively. These friction angles of MICP-treated sands ilar to those reported by Martinez et al. (2013). Martinez et al.
are similar to those of untreated sands. However, the cohesion (2013) reported that the CaCO3 content profiles are generally
increased after MICP treatment, showing residual cohesion at determined by the microbe distribution along the specimen.

© ASCE 04015066-5 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2016, 142(2): 04015066


500 500
50/70 sand 20/30 sand
Effective confining
400 pressure (kPa) 400 Effective confining

Deviator stress (kPa)

Deviator stress (kPa)


pressure (kPa)

300 300
100
100
200 200
50 50
100 25 100
25

0 0
0 2 4 6 8 10 12 0 2 4 6 8 10 12
(a) Axial strain (%) (b)
Downloaded from ascelibrary.org by Technische Universiteit Delft on 03/28/24. Copyright ASCE. For personal use only; all rights reserved.

Axial strain (%)

2.5 2.5
50/70 sand 20/30 sand
2.0 2.0
Volumetric strain (%)

Volumetric strain (%)


1.5 1.5

1.0 1.0

0.5 0.5

Dilation Dilation
0.0 0.0
Contraction Contraction
-0-.5 -0-.5
0 2 4 6 8 10 12 0 2 4 6 8 10 12
(c) Axial strain (%) (d) Axial strain (%)
MICP 25
MICP 50
MICP 100
Untreated 25
Untreated 50
Untreated 100

Fig. 4. Consolidated drained triaxial test results: (a) stress-strain for 50=70 sand; (b) stress-strain for 20=30 sand; (c) volumetric strain for 50=70
sand; (d) volumetric strain for 20=30 sand

Furthermore, the distribution of CaCO3 along the 50=70 sand triaxial tests presented in this paper show an increase of the soil
specimens may be affected by bioclogging (i.e., a decrease in void strength in the 20=30 sand at 1% calcium carbonate content.
ratio and hydraulic conductivity near the solution injection and
effluent points owing to CaCO3 precipitation). In contrast, the
CaCO3 content of the coarser, more permeable 20=30 sand has Sands Treated with 0.3 M CaCl2
a uniform distribution along the height of the specimens. As discussed in the “Background” section, Lin et al. (2015), who
It is also worth noting that the 0.1 M CaCl2 =MICP-treated investigated the effects of MICP treatment on the static axial capac-
20=30 sand specimens have lower average CaCO3 content (1%) ity and soil-pile interaction of pervious concrete piles, reported that
compared with the 50=70 sand specimens (1.6%), which were the CaCO3 content surrounding the tested piles ranged from 0 to
treated using the same procedure and solution. The average in- 3%. Therefore, the effects of CaCO3 cementation levels on the
crease in S-wave velocity, peak deviator stress and cohesion for 50=70 sand were investigated using a cementation medium with
the 0.1 M CaCl2 =MICP-treated 20=30 sand (2.6 times, 171%, 0.3 M CaCl2 to target a CaCO3 content closer to 3%. Fig. 6(a)
and 58 kPa, respectively) are higher than that of the 50/70 sand shows the average CaCO3 content of the triaxial specimens treated
(1.8 times, 93% and 41 kPa, respectively). This demonstrates that using 0.3 M CaCl2 cementation medium. Using the same treatment
the mechanical properties of MICP-treated sands are not only con- procedure (five flushes of the cementation medium), the average
trolled by the average or bulk CaCO3 content. The spatial distri- CaCO3 content of the specimens treated using the 0.3 M CaCl2
bution of CaCO3 in the pore space (e.g., the effective CaCO3 concentration was 2.5%, which is higher than the 1.6% for the
content at particle contacts), which is affected by factors such as specimens treated with the 0.1 M CaCl2 solution. The calculated
particles size, pore size, particle surface area, and bacteria distribu- yield of CaCO3 (weight of precipitated CaCO3 in the specimen/
tion, is an important factor controlling the properties of MICP- equivalent weight of CaCO3 from the input cementation media)
treated sands (Rebata-Landa 2007; Cheng et al. 2013; Martinez were 57 and 29% for the 50=70 sand specimens treated by 0.1
et al. 2013). It is also worth noting that unlike Whiffin et al. and 0.3 M CaCl2 , respectively, and 37% for 20=30 sand specimen
(2007), who reported a threshold calcium carbonate content of treated by 0.1 M CaCl2 . To acquire a high yield efficiency (>80%),
approximately 3.6% by weight before the unconfined compressive Al Qabany et al. (2012) reported that the threshold input rate of urea
strength of MICP-treated sand started to increase, the results of and CaCl2 should be slow (lower than 0.042 mol=L=h). However,

© ASCE 04015066-6 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2016, 142(2): 04015066


1000
n
0
5070 untreated Ei=kPa(σ3/Pa) from Janbu (1963) Sample
5070 MICP Confining pressure (kPa) top
2030 untreated 20
25
2030 MICP k=2478.8
n=0.09 50
40 100

Sample depth (mm)


25 0.3 M
Ei (MPa)

k=1703.6
100
n=0.19
60
Medium
k=636.7 80 injection
n=0.53 direction

k=641.9 100
n=0.88 Sample
Downloaded from ascelibrary.org by Technische Universiteit Delft on 03/28/24. Copyright ASCE. For personal use only; all rights reserved.

10 120 bottom
10 1 00 Average
(a) CaCO3 content 1.5% 1.6% 1.6% 2.5%
Effective confining pressure (kPa)
140
700 0 1 2 3 4 5
160
Untreated 25
155 (a) CaCO3 content (CaCO3 (g)/sand (g), %)
50/70 sand
600 150
MICP 25
145 MICP 50 0
Lines connected to points Sample
S-wave velocity (m/s)

140 MICP 100


500 when axial strain=
0.0010.01 0.1 01 0%
010 0100 Untreated 25
top
Untreated 50
20 Confining pressure (kPa)
Untreated 100
400 25
40

Sample depth (mm)


50
100
300
60
Medium
200 injection
80 direction
100
0.001 0.01 0.1 1 10 100 100
(b) Axial strain (%) Sample
120 bottom
700 Average
Lines connected to points when axial strain= 0% 0.9% 1.1% 1.0% CaCO3 content
140
600 0 1 2 3 4 5
(b) CaCO3 content (CaCO3 (g)/sand (g), %)
S-wave velocity (m/s)

500 20/30 sand


MICP 25
MICP 50 Fig. 6. CaCO3 content distribution of (a) 50=70 sand; (b) 20=30 sand
400 MICP 100 for specimens treated with 0.1 M concentration of CaCl2 except where
Untreated 25 noted in the legend
Untreated 50
300 Untreated 100

200
(Al Qabany et al. 2011; Weil et al. 2011). The S-wave velocities
versus CaCO3 content reported herein are in a good agreement with
100
0.001 0.01 0.1 1 10 100 the results reported in the literature (Al Qabany et al. 2011; Weil
(c) Axial strain (%) et al. 2011). P-wave velocities of the specimens treated with 0.3 M
CaCl2 increased by an average of 10.1% relative to untreated spec-
Fig. 5. Stiffness during shear loading for specimens treated with 0.1 M imens, which is compared with 8.7% increase for specimens treated
CaCl2 cementation medium: (a) initial tangent modulus; (b) S-wave with 0.1 M CaCl2 solution.
velocity of 50=70 sand; (c) S-wave velocity of 20/30 sand (numbers Fig. 7(a) compares the stress-strain responses of the 50=70 sand
in the legend are effective confining pressures in kPa) at a confining pressure of 25 kPa for specimens treated with 0.1 M
and 0.3 M CaCl2 solutions. At this confining pressure, the peak
deviator stress of the specimens treated with 0.3 M CaCl2 increased
the urea and CaCl2 input rates in our tests were 0.27 and by an average factor of 4.8 relative to the untreated specimens,
0.8 mol=L=h for specimens treated with 0.1 and 0.3 M CaCl2 , re- whereas the peak deviator stress of the specimens treated with
spectively, which were controlled by injecting two pore volumes of 0.1 M CaCl2 solution increased by a factor of 1.4. Furthermore,
solution over an interval of 3 h. the specimens treated with the 0.3 M CaCl2 solution showed a
During treatment, the S-wave velocity of the specimens treated 45% increase in the deviator stress at an axial strain of 10% com-
with 0.3 M CaCl2 show an average increase by a factor of 5.5 pared with the untreated specimens, whereas the specimens treated
compared with the untreated specimens, which is compared with the 0.1 M CaCl2 solution had essentially the same deviator
with a factor of 1.8 for specimens treated with 0.1 M CaCl2 stress at a strain of 10% compared with the untreated specimens.
solution. It is worth mentioning, however, that there is no direct The higher ultimate deviator stress of the specimens treated with
relationship between the S-wave velocities and CaCO3 content 0.3 M CaCl2 could be attributed to the increased particle roughness

© ASCE 04015066-7 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2016, 142(2): 04015066


Downloaded from ascelibrary.org by Technische Universiteit Delft on 03/28/24. Copyright ASCE. For personal use only; all rights reserved.

Fig. 7. Effect of CaCl2 concentration of cementation medium on consolidated drained triaxial test behavior of 50=70 sand at 25 kPa effective
confining pressure: (a) stress-strain behavior; (b) volumetric strain; (c) S-wave velocity during shear loading (numbers in the legend are CaCl2
concentrations)

resulting from the precipitated CaCO3 (Montoya and DeJong 2015) concentrated shear band failure, as shown in Fig. 7(b). The change
and residual cohesion in the specimens (Clough et al. 1981). of volumetric strain behavior may be controlled by the formation of
The volumetric strain versus axial strain data presented in the shear band. The shear band was more concentrated in the spec-
Fig. 7(b) shows that the 50/70 specimens treated with MICP were imens with CaCO3 content of 2.5% (0.3 M CaCl2 ) than that with
more dilatant than the untreated specimens. It is worth noting that CaCO3 content of 1.6% (0.1 M CaCl2 ). At small strain, more di-
untreated specimens showed bulging failure, whereas, as CaCO3 lation was experienced by the specimens with CaCO3 content of
content increases, the specimens treated by MICP manifest more 2.5% (0.3 M CaCl2 ) than by the specimens with CaCO3 content

© ASCE 04015066-8 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2016, 142(2): 04015066


0.74 0.74
Average CaCO3 content 20/30 sand
0.72 0.72 Untreated
0.70 0.70 MICP 0.1 M
0% MICP 0.3 M
0.68 0.68 Average CaCO3 content

Void ratio, e
Void ratio, e
1.4%
0.66 2.6% 0.66
0.64 0.64
50/70 sand
0.62 0.62 0%
Untreated
0.60 MICP 0.1 M 0.60 0.6%

0.58 MICP 0.3 M 0.58 1.6%

0.56 0.56
10 100 1000 10000 10 100 1000 10000
(a) Vertical effective stress (kPa) (b) Vertical effective stress (kPa)
Downloaded from ascelibrary.org by Technische Universiteit Delft on 03/28/24. Copyright ASCE. For personal use only; all rights reserved.

Fig. 8. Confined compression test results for 50=70 and 20=30 sands: (a) compressibility for 50=70 sand; (b) compressibility for 20=30 sand

of 1.6% (0.1 M CaCl2 ). As the strain increased, the soil particles compression tests were lower (with a difference ranging from
movement is localized to the shear band, resulting in smaller vol- 0.2 to 0.4%) than that measured from triaxial tests using the same
ume change than that of the specimens with 0.1 M CaCl2 . Similar treatment procedure. This difference could be attributed to the pore
shear band formation at high CaCO3 cementation level was also size distribution, bacteria distribution and urease activity along the
recently reported by Montoya and DeJong (2015). sample, and their influence on CaCO3 distribution (Rebata-Landa
Using the measured S-wave and P-wave velocities, the 2007; Cheng and Cord-Ruwisch 2014; Martinez et al. 2013). In
Poisson’s ratio of the specimens can be calculated using Eq. (4) addition, the observed higher cementation levels around the inlets
rffiffiffiffiffi rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi or outlets of the triaxial specimens could be another reason for this
Vp M 2ð1 − υÞ difference.
¼ ¼ ð4Þ
Vs G 1 − 2υ Figs. 8(a and b) show the variation of void ratio as a function of
vertical effective stresses for treated and untreated specimens of
where V p and V s are the P-wave and S-wave velocities, G and M both sands. The MICP-treated specimens are less compressible
are shear and constrained moduli, and υ is the Poisson’s ratio. The compared with the untreated specimens, and the specimens with
initial (before treatment) Poisson’s ratio was 0.496. After MICP higher CaCO3 content are less compressible than the specimens
treatment, the Poisson’s ratio decreased to 0.467 and 0.435 for of the same grain size with lower CaCO3 content. The results of
the 50=70 sand treated by 0.1 M and 0.3 M CaCl2 , respectively. void ratio versus vertical effective stresses are in good agreement
The Poisson’s ratios at the end of the testing (axial strain with the results reported by Feng and Montoya (2014). The calcu-
∼10%) were 0.486 and 0.440 for the 50=70 sand treated by 0.1 lated compression index (Cc ) of 50=70 sand decreased from 0.024
M and 0.3 M CaCl2 , respectively. These results suggest that the
(0% CaCO3 content) to 0.009 (2.6% CaCO3 content). The calcu-
specimens treated with 0.3 M CaCl2 experienced less lateral
lated compression indexes (Cc ) of 20=30 sand decreased from
deformation (e.g., less volume change at the same axial strain) than
0.019 (0% CaCO3 content) to 0.009 (1.6% CaCO3 content). This
that of the specimens treated with 0.1 M CaCl2 .
decrease of compression index is similar to that of residual soils
During triaxial compression loading, a similar degradation of
treated by MICP before the fracture of CaCO3 bonds presented
S-wave velocity was observed in the specimens treated with 0.3
by Lee et al. (2013).
M CaCl2 , compared with the specimen treated with 0.1 M
Figs. 9(a–d), present the variation of normalized S-wave veloc-
CaCl2 , as shown in Fig. 7(c). However, the S-wave velocity at
ities (S-wave velocity divided by initial S-wave velocity) for the
an axial strain of 10% was still substantially higher in the speci-
horizontal and vertical directions as a function of vertical effective
mens treated with 0.3 M CaCl2 solution compared with either
stress (σv0 ). The untreated specimens show confined compression
the specimens treated with 0.1 M CaCl2 solution or the untreated
behavior characteristic of sand, with the normalized S-wave veloc-
specimens. Higher S-wave velocity at 10% axial strain in the spec-
ities increasing as the vertical effective stress increases and then
imens treated with 0.3 M CaCl2 could be primarily attributed to the
nonfractured bonding between soil particles away from the concen- decreasing during unloading, though with some hysteresis (i.e., nor-
trated shear band shown by the pictures in Fig. 7(b). malized S-wave velocities are higher after unloading than during
initial (virgin) loading for the same vertical effective stress). The
MICP-treated specimens have higher initial normalized S-wave
Confined Compression Tests velocities than that of untreated specimens and show a different
pattern. For the specimens treated with 0.1 M CaCl2 and thus with
Confined compression tests were also performed on untreated and the smaller amounts of CaCO3 content (the 50=70 sand with 1.4%
MICP-treated sands. S-wave velocities in both the vertical direction and the 20=30 sand with 0.6%), there was initially no hysteresis and
(V v ) and the horizontal direction (V vh ) were measured. The CaCO3 then a lower shear wave velocity at the same effective stress during
content of the MICP-treated confined compression specimens was unloading as in the loading phase (trending towards the same value
measured in a similar manner as for the triaxial test specimens. The as for the untreated specimen after unloading). For the specimens
CaCO3 content of the 50/70 sand was 1.4% when treated with 0.1 treated with 0.3 M CaCl2 and thus with greater amounts of CaCO3
M CaCl2 solution and 2.6% when treated with 0.3 M CaCl2 sol- (the 50=70 sand with 2.6% and the 20=30 sand with 1.6%), there
ution. The CaCO3 content of the 20=30 sand was 0.6% when was no hysteresis for the shear wave velocity in the horizontal di-
treated with 0.1 M CaCl2 solution and 1.6% when treated with rection and less hysteresis for the vertical direction shear wave
0.3 M CaCl2 solution (Table 3). The comparison between Tables 2 velocity during unloading compared with the shear wave velocity
and 3 shows that the CaCO3 content measured from confined of specimens treated with 0.1 M CaCl2 solution.

© ASCE 04015066-9 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2016, 142(2): 04015066


Normalized horizontal S-wave velocity (Vhv/Vo)

Normalized horizontal S-wave velocity (Vhv/Vo)


10 10
Average CaCO3 content
2.6%
1.6%
Note: V0=85 m/s
Loading 1.4% 0.6%
Unloading 0%
Average CaCO3 0%
content Note: V0=100 m/s
Loading
50/70 sand Unloading 20/30 sand
Untreated Vhv Untreated Vhv
MICP 0.1 M Vhv MICP 0.1 M Vhv
MICP 0.3 M Vhv MICP 0.3 M Vhv
1 1
10 100 1000 10000 10 100 1000 10000
Downloaded from ascelibrary.org by Technische Universiteit Delft on 03/28/24. Copyright ASCE. For personal use only; all rights reserved.

Vertical effective stress (kPa) Vertical effective stress (kPa)


(a) (b)

Normalized vertical S-wave velocity (Vv/Vo)


Normalized vertical S-wave velocity (Vv/Vo)

10 10
2.6% Average CaCO3 content

1.6%
Note: V0=72 m/s
Loading 1.4%
0.6%
Unloading 0%
0%
Average CaCO3
Note: V0= 94 m/s
content
Loading
50/70 sand Unloading
Untreated Vv Untreated Vv
MICP 0.1 M Vv MICP 0.1 M Vv
MICP 0.3 M Vv 20/30 sand
MICP 0.3 M Vv
1 1
10 100 1000 10000 10 100 1000 10000

(c) Vertical effective stress (kPa) (d) Vertical effective stress (kPa)

Fig. 9. Confined compression test results for 50=70 and 20=30 sands: (a) V hv for 50=70 sand; (b) V hv for 20=30 sand; (c) V v for 50=70 sand;
(d) V v for 20=30 sand [V hv is the S-wave velocity transmitted in horizontal direction, and V v is the S-wave velocity transmitted in vertical
direction as show in Fig. 2(c)]

The measured S-wave velocity as a function of vertical effective M CaCl2 treatment (1.6% CaCO3 content for 50=70 sand and 1%
stress for both MICP-treated sands were compared with the CaCO3 content for 20=30 sand) show CaCO3 precipitated at par-
loose Nevada sand modified using portland cemented (Yun and ticle contacts and coating particle surfaces. As the CaCl2 treatment
Santamarina 2005). The comparison shows similar trends of concentration increases from 0.1 M and 0.3 M, the images of the
S-wave velocities as the effective stress increases. However, the MICP-treated 50=70 sands show an increase in the CaCO3 content
sudden change of S-wave velocity as the stress increases in samples (area increased shown in EDS Ca element mapping) consistent with
treated with portland cement was not observed in MICP-treated the CaCO3 content measurement (CaCO3 content increased from
sand specimens. This could be attributed to the higher void ratio 1.6 to 2.5%).
(eo ¼ 1.1) of the loose Nevada sand, which may induce collapse Based on the shape of the CaCO3 crystals in Figs. 11(a–c), two
during loading as discussed by Feng and Montoya (2014). Further- types of CaCO3 morphologies, calcite and vaterite crystals, are
more, small-strain shear modulus can be calculated using the present in both the 50=70 and 20=30 sands. In the 20=30 sand,
measured S-wave velocity Eq. (5) smaller sand particles (∼400 μm) can be seen in Figs. 11(b and c)
filling the pore space created by larger particles. Figs. 11(d–f) show
G ¼ ρsat × v2s ð5Þ
that all three of the idealized CaCO3 distributions shown in Fig. 1
where ρsat = saturated density of sand. The calculated small-strain have occurred in the treated specimens, i.e., CaCO3 can be seen to be
shear modulus versus vertical effective stress on log-scale for deposited at particle contacts [contact cementing, Fig. 1(a)], coating
50=70 sands is similar to that reported by Montoya et al. (2013) particles [grain coating, Fig. 1(b)], and growing into the pore space
for untreated, and MICP moderately-treated specimens with to create cement bridges in the sand matrix [matrix supporting,
CaCO3 content of ∼2.6%. Fig. 1(c)].

Microscale Properties of MICP-Treated Sands Conclusions

The SEM and EDS images from the triaxial test specimens are This paper describes an investigation into the mechanical behavior
shown in Fig. 10. To investigate a similar range of CaCO3 content of MICP-treated Ottawa 20=30 and 50=70 silica sands using
(0 to ∼3%) reported by Lin et al. (2015), the specimens treated by drained triaxial and confined compression tests with P-wave and
0.3 M CaCL2 was also investigated by SEM and EDS images. The S-wave velocity measurements. Data on the microscale CaCO3
images of MICP-treated 50=70 and 20=30 samples subjected to 0.1 distribution from SEM and EDS images are also presented herein.

© ASCE 04015066-10 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2016, 142(2): 04015066


Downloaded from ascelibrary.org by Technische Universiteit Delft on 03/28/24. Copyright ASCE. For personal use only; all rights reserved.

Fig. 10. Scanning electron microscopy (SEM) and energy-dispersive X-ray spectroscopy (EDS) images of untreated and MICP-treated samples for
50=70 and 20=30 sands

The tests results reported in this paper will be employed to further CaCO3 content, the S-wave velocity increases as the
investigate the use of MICP modification and the effects of CaCO3 CaCO3 content increases. During loading in triaxial tests,
content on improving the static axial capacity and soil-pile inter- S-wave velocity shows degradation as the axial strain in-
action of permeable piles. Based on the data presented in this paper, creases. Combining measured S-wave and P-wave velocities,
the following observations were made and conclusions were drawn: the Poisson’s ratio was calculated and shows a decrease as
1. The S-wave velocity was used to monitor MICP cementation CaCO3 content increases;
during treatment and loading phases. Although no direct rela- 2. During triaxial tests, the peak deviator stress of the Ottawa
tionship was established between the S-wave velocity and 20=30 sand with CaCO3 content of 1.6% and the Ottawa

Fig. 11. SEM images showing CaCO3 morphologies and spatial distributions of CaCO3 around sand surface

© ASCE 04015066-11 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2016, 142(2): 04015066


50=70 sand with CaCO3 content of 1% increased by an average Burbank, M., Weaver, T., Lewis, R., Williams, T., Williams, B., and
of 93 and 171%, respectively, compared with their correspond- Crawford, R. (2013). “Geotechnical tests of sands following bioinduced
ing untreated specimens. MICP treatment can be seen to signif- calcite precipitation catalyzed by indigenous bacteria.” J. Geotech. Geo-
icantly increase the dilatancy of both sands. Observation of the environ. Eng., 10.1061/(ASCE)GT.1943-5606.0000781, 928–936.
failure mode reveals bulging failures for the untreated speci- Chaney, R. C., and Mulilis, J. P. (1978). “Suggested method for soil speci-
men remolding by wet-raining.” Geotech. Test. J., 1(2), 107–110.
mens. As CaCO3 content increases, the peak strength increases,
Cheng, L., and Cord-Ruwisch, R. (2014). “Upscaling effects of soil im-
and more concentrated shear band has formed, resulting in high-
provement by microbially induced calcite precipitation by surface per-
er dilation at small strain and lower dilation at large strains; colation.” Geomicrobiol. J., 31(5), 396–406.
3. Triaxial results show that initial tangent Young’s moduli (Ei ) Cheng, L., Cord-Ruwisch, R., and Shahin, M. A. (2013). “Cementation of
of MICP-treated sands are controlled by the CaCO3 content sand soil by microbially induced calcite precipitation at various degrees
and are less sensitive to the increase of the effective confining of saturation.” Can. Geotech. J., 50(1), 81–90.
pressure than untreated specimens. In confined compression Chou, C., Seagren, E. A., Aydilek, A. H., and Lai, M. (2011). “Biocalci-
tests, as the CaCO3 content increases, the stiffness of the speci- fication of sand through ureolysis.” J. Geotech. Geoenviron. Eng.,
Downloaded from ascelibrary.org by Technische Universiteit Delft on 03/28/24. Copyright ASCE. For personal use only; all rights reserved.

mens increases and becomes less sensitive to the increase in 10.1061/(ASCE)GT.1943-5606.0000532, 1179–1189.
normal stress; Clayton, C. R. I. (2011). “Stiffness at small strain: Research and practice.”
4. Although the 20=30 sand had a lower CaCO3 content com- Géotechnique, 61(1), 5–37.
pared with 50=70 sand when treated using a 0.1 M CaCl2 so- Clough, G. W., Rad, N. S., Bachus, R. C., and Sitar, N. (1981). “Cemented
lution, the increase of S-wave velocity, peak shear strength, sands under static loading.” J. Geotech. Eng. Div., 107(6), 799–817.
and cohesion for the MICP-treated 20=30 sand were higher than DeJong, J. T., et al. (2010a). “Soil engineering in vivo: Harnessing natural
biogeochemical systems for sustainable, multi-functioning engineering
that of the MICP-treated 50=70 sand. This indicates that the
solutions.” J. R. Soc. Interface, rsif20100270.
strength and stiffness of MICP-treated sands are controlled by
DeJong, J. T., et al. (2013). “Biogeochemical processes and geotechnical
factors other than the average or bulk CaCO3 content of the speci- applications: Progress, opportunities and challenges.” Géotechnique,
mens, e.g., by the effective CaCO3 content at particle contacts; 63(4), 287–301.
5. The MICP-treated specimens are less compressible than un- DeJong, J., Fritzges, M., and Nüsslein, K. (2006). “Microbially induced
treated specimens. As the CaCO3 content increases, the com- cementation to control sand response to undrained shear.” J. Geotech.
pressibility of the treated soil specimens decreases. As CaCO3 Geoenviron. Eng., 10.1061/(ASCE)1090-0241(2006)132:11(1381),
content increases, the S-wave velocity shows less decrease or 1381–1392.
no change as the normal stress increases, and there is less hys- DeJong, J. T., Mortensen, B. M., Martinez, B. C., and Nelson, D. C.
teresis for the S-wave velocity upon unloading; and (2010b). “Bio-mediated soil improvement.” Ecol. Eng., 36(2), 197–210.
6. SEM and EDS images show that the three idealized CaCO3 Ebigbo, A., et al. (2012). “Darcy-scale modeling of microbially induced
distributions can occur in the treated specimens, i.e., CaCO3 carbonate mineral precipitation in sand columns.” Water Resour.
can deposit at particle contacts (contact cementing), coat par- Res., 48(7).
ticles (grain coating), and grow into the pore space (matrix Feng, K., and Montoya, B. M. (2014). “Behavior of bio-mediated soil in k0
loading.” New Frontiers in Geotechnical Engineering Geo-Shanghai
supporting).
Proc., ASCE, Reston, VA, 1–10.
Ferris, F. G., Stehmeier, L. G., Kantzas, A., and Mourits, F. M. (1996).
“Bacteriogenic mineral plugging.” J. Can. Pet. Technol., 35(8), 56–61.
Acknowledgments Fujita, Y., Taylor, J. L., Wendt, L. M., Reed, D. W., and Smith, R. W.
(2010). “Evaluating the potential of native ureolytic microbes to re-
The authors acknowledge the support of the Civil, Mechanical, and mediate a 90Sr contaminated environment.” Environ. Sci. Technol.,
Manufacturing Innovation (CMMI) Division at National Science 44(19), 7652–7658.
Foundation (No. 1233566). The research team acknowledges the Hamdan, N., Kavazanjian, E., Jr., Rittmann, B. E., and Karatas, I. (2011).
efforts of several undergraduate students, including Pierre Bick, “Carbonate mineral precipitation for soil improvement through micro-
Alexa Hendricks, and Yassira Alaziz, and graduate students includ- bial denitrification.” Proc., GeoFrontiers 2011: Advances in Geotech-
ing Hang Dong, Hankai Zhu, Suguang Xiao, and Lusu Ni. Also, the nical Engineering, ASCE, Reston, VA, 3925–3934.
authors acknowledge the help of Dan Zeroka and Edward Tomlin- Ivanov, V., and Chu, J. (2008). “Applications of microorganisms to geo-
son, technician, and instrumentation and system specialist, respec- technical engineering for bioclogging and biocementation of soil in
tively, at Lehigh University’s Advanced Technology for Large situ.” Rev. Environ. Sci. BioTechnol., 7(2), 139–153.
Structural Systems (ATLSS) Engineering Research Center. Janbu, N. (1963). “Soil compressibility as determined by oedometer and
triaxial tests.” Proc., 3rd European Conf. on Soil Mechanics and Foun-
dation Engineering, Vol. 1, Wiesbaden, Germany, 19–25.
Jones, D., Pell, P. A., and Sneath, P. H. A. (1991). “Maintenance of bacteria
References
on glass beads at −60°C to −76°C.” Maintenance of microorganisms
Al Qabany, A., Mortensen, B., Martinez, B., Soga, K., and DeJong, J. and cultured cells: A manual of laboratory methods, B. E. Kirsop and
(2011). “Microbial carbonate precipitation: Correlation of S-wave A. Doyle, eds., Academic Press, San Diego, 45–50.
velocity with calcite precipitation.” Proc., ASCE Geo-Frontiers 2011 Karatas, I., Kavazanjian, E., Jr., and Rittmann, B. E. (2008). “Microbially
Conf., ASCE, Reston, VA. induced precipitation of calcite using Pseudomonas denitrificans.”
Al Qabany, A., and Soga, K. (2013). “Effect of chemical treatment Proc., 1st Int. Conf. on Biogeotechnical Engineering, Technical Univ.
used in MICP on engineering properties of cemented soils.” Géotech- of Delft, Delft, Netherlands.
nique, 63(4), 331–339. Kavazanjian, E., Jr., Iglesias, E., and Karatas, I. (2009). “Biopolymer
Al Qabany, A., Soga, K., and Santamarina, C. (2012). “Factors affecting soil stabilization for wind erosion control.” Proc., 17th Int. Conf. on
efficiency of microbially induced calcite precipitation.” J. Geotech. Geo- Soil Mechanics and Geotechnical Engineering, ISSMGE, London,
environ. Eng., 10.1061/(ASCE)GT.1943-5606.0000666, 992–1001. 881–884.
Bang, S. S., Bang, S., Frutiger, S., Nehl, L. M., and Comes, B. L. (2009). Kavazanjian, E., Jr., and Karatas, I. (2008). “Microbiological improvement
“Application of novel biological technique in dust suppression.” Trans- of the physical properties of soil.” Symp. to Honor James K. Mitchell,
portation Research Board 88th Annual Meeting, TRB, Washington, Proc., 6th Int. Conf. on Case Histories in Geotechnical Engineering,
DC. Missouri Univ. of Science and Technology, Rolla, MO.

© ASCE 04015066-12 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2016, 142(2): 04015066


Lee, J., and Santamarina, J. C. (2005). “Bender elements: Performance and Stocks-Fischer, S., Galinat, J. K., and Bang, S. S. (1999). “Microbiological
signal interpretation.” J. Geotech. Geoenviron. Eng., 10.1061/(ASCE) precipitation of CaCO3.” Soil Biol. Biochem., 31(11), 1563–1571.
1090-0241(2005)131:9(1063), 1063–1070. Tatsuoka, F., and Shibuya, S. (1991). “Deformation characteristics of soils
Lee, M. L., Ng, W. S., and Tanaka, Y. (2013). “Stress-deformation and and rocks from field and laboratory tests.” 9th Asian Regional Conf. on
compressibility responses of bio-mediated residual soils.” Ecol. Eng., Soil Mechanics and Foundation Engineering, ISSMGE, London.
60, 142–149. van der Ruyt, M., and van der Zon, W. (2009). “Biological in situ reinforce-
Lin, H., Suleiman, M., Jabbour, H., and Brown, D. (2015). “Enhancement ment of sand in near-shore areas.” Proc. ICE—Geotech. Eng., 162(1),
of pervious concrete pile subjected to uplift load using microbial in- 81–83.
duced carbonate precipitation.” IFCEE 2015, ASCE, Reston, VA, van Paassen, L. A. (2009). “Biogrout ground improvement by microbially
775–783. induced carbonate precipitation.” Ph.D. thesis, Delft Univ. of Technol-
Martinez, B. C., et al. (2013). “Experimental optimization of microbial- ogy, Delft, Netherlands.
induced carbonate precipitation for soil improvement.” J. Geotech. van Paassen, L. A., van Loosdrecht, M. C. M., Pieron, M., Mulder, A.,
Geoenviron. Eng., 10.1061/(ASCE)GT.1943-5606.0000787, 587–598. Ngan-Tillard, D. J. M., and van der Linden, T. J. M. (2012). “Strength
and deformation of biologically cemented sandstone.” Rock engineer-
Martinez, B. C., and DeJong, J. T. (2009). “Bio-mediated soil improve-
Downloaded from ascelibrary.org by Technische Universiteit Delft on 03/28/24. Copyright ASCE. For personal use only; all rights reserved.

ing in difficult ground conditions, soft rocks and karst, Taylor & Francis
ment: Load transfer mechanisms at the micro- and macro-scales.”
Group, London, 405–410.
U.S.-China Workshop on Ground Improvement Technologies, ASCE,
Weil, M. H., DeJong, J. T., Martinez, B. C., and Mortensen, B. M. (2011).
Reston, VA, 242–251.
“Seismic and resistivity measurements for real-time monitoring of mi-
Mitchell, J. K., and Santamarina, J. C. (2005). “Biological considerations in crobially induced calcite precipitation in sand.” Geotech. Test. J., 35(2),
geotechnical engineering.” J. Geotech. Geoenviron. Eng., 10.1061/ 330–341.
(ASCE)1090-0241(2005)131:10(1222), 1222–1233. Whiffin, V. S., van Paassen, L. A., and Harkes, M. P. (2007). “Microbial
Montoya, B. M., and Dejong, J. T. (2015). “Stress-strain behavior of sands carbonate precipitation as a soil improvement technique.” Geomicro-
cemented by microbially induced calcite precipitation.” J. Geotech. biol. J., 24(5), 417–423.
Geoenviron. Eng, 141(6), 04015019. Yun, T. S., and Santamarina, J. C. (2005). “Decementation, softening, and
Montoya, B. M., Dejong, J. T., and Boulanger, R. W. (2013). “Dynamic collapse: Changes in small-strain shear stiffness in Ko loading.” J. Geo-
response of liquefiable sand improved by microbial-induced calcite tech. Geoenviron. Eng., 10.1061/(ASCE)1090-0241(2005)131:3(350),
precipitation.” Géotechnique, 63(4), 302–312. 350–358.
Rebata-Landa, V. (2007). “Microbial activity in sediments: Effects Yun, T. S., Santamarina, J. C., and Ruppel, C. (2007). “Mechanical proper-
on soil behavior.” Ph.D. dissertation, Georgia Institute of Technology, ties of sand, silt, and clay containing tetrahydrofuran hydrate.” J.
Atlanta. Geophys. Res. Solid Earth, 112(B4).

© ASCE 04015066-13 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2016, 142(2): 04015066

You might also like