Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Key Factors for Deep Cement Mixing Construction

for Undredged Offshore Land Reclamation


K. S. Yin 1; L. M. Zhang, F.ASCE 2; H. F. Zou 3; H. Y. Luo 4; and W. J. Lu 5
Downloaded from ascelibrary.org by Hong Kong University of Sci and Tech (HKUST) on 11/07/22. Copyright ASCE. For personal use only; all rights reserved.

Abstract: Deep cement mixing (DCM) is an environmentally friendly technique for offshore ground improvement without dredging or
much disturbance to the marine ecological system. Several field construction factors can influence the unconfined compressive strength
(UCS) of cement-stabilized soil. In this study, key construction factors are evaluated referring to site investigation records, construction
records, and quality test results of a large offshore DCM construction project. The key factors were attributed to geological conditions,
construction procedures, and curing conditions. Specific field construction factors include fluctuations of tidal level, original soil type, vol-
ume fraction of injected water, volume fraction of injected cement slurry, injection rate of cement slurry, penetrating and mixing time per
meter, curing age, and moisture content. The importance of these construction factors on the DCM strength has been quantified using a
statistical method based on construction records. The injected water volume and original soil type are noted to be the two most dominant
factors on the UCS of the treated soils. Longer mixing time improved the strength of the treated soils. DOI: 10.1061/(ASCE)GT.1943-
5606.0002848. © 2022 American Society of Civil Engineers.
Author keywords: Deep cement mixing (DCM); Ground improvement; Land reclamation; Offshore construction; Unconfined compressive
strength (UCS); Marine deposit.

Introduction deep cement mixing (DCM) widely used in offshore undredged


land reclamations. DCM improves the consistency, shear strength,
Land reclamation is one way to ease the land shortage problem in deformation characteristics, and permeability of natural marine
coastal areas. Deep mixing is one of the most preferred techniques soils for practical engineering purposes, including bearing capacity
for undredged land reclamation due to its capability to improve the improvement, settlement reduction, seepage control, and excava-
strength and stiffness of soft marine clays and avoid up-heave prob- tion support. Famous land reclamation projects involving large-
lems. This technique also minimizes disturbance to the natural scale DCM include the second-phase artificial island of Kansai
marine ecological system from dredging activities. The deep mix- International Airport (Furudoi 2005) and the D-runway project
ing technique was originated in Japan in the late 1960s and adopted of Haneda International Airport (Watabe and Noguchi 2011).
worldwide starting in the 1970s. The injected binders, either in the Laboratory tests and numerical analyses have been performed to
form of dry powder or wet slurry containing cement or lime, are evaluate the mechanical characteristics of DCM-improved soil con-
mixed with in-situ soils during construction, forming columns or sidering numerous factors, including water content (Horpibulsuk
clusters of cemented binder–soil mixtures. This study concerns et al. 2012), stress concentration ratio (Jiang et al. 2013), column
diameter, length and distribution (Liu et al. 2012), column stiffness
1
Postdoctoral Research Associate, Dept. of Civil and Environmental (Yapage et al. 2014), column material (Abusharar et al. 2009), col-
Engineering, Hong Kong Univ. of Science and Technology (HKUST), umn permeability (Yin and Fang 2006), surrounding soil properties
Kowloon, Hong Kong 999077, China. ORCID: https://orcid.org/0000 (Rogers and Glendinning 1997), and penetration ratio (Yang et al.
-0003-1350-4497 2014). Other studies investigated key contributory factors to DCM
2
Chair Professor and Head, Dept. of Civil and Environmental Engineer- mixture strength, including soil type, water content, binder amount
ing, Hong Kong Univ. of Science and Technology, Kowloon, Hong Kong and type, curing period, mixing time, and humidity (Kitazume
999077, China; Professor, HKUST Shenzhen-Hong Kong Collaborative et al. 2015; Mohammadinia et al. 2019; Disfani et al. 2021;
Innovation Research Institute, Shenzhen 518000, China (corresponding
Subramaniam and Banerjee 2020; Bellato et al. 2020). The uncon-
author). ORCID: https://orcid.org/0000-0001-7208-5515. Email: cezhang@
ust.hk
fined compressive strength (UCS) of cement–soil mixtures exhib-
3
Assistant Engineer, AECOM Asia Limited Company, 138 Shatin Rural ited significant spatial variation (Lee et al. 2006; Chen et al. 2011;
Committee Rd., Hong Kong 999077, China. Wijerathna and Liyanapathirana 2018, 2019), with measured auto-
4
Postdoctoral Research Associate, Dept. of Civil and Environmental correlation distances ranging from far less than a cluster’s diameter
Engineering, Hong Kong Univ. of Science and Technology, Kowloon, to around 50 m (Honjo 1982; Larsson et al. 2005a, b; Navin 2005),
Hong Kong 999077, China. ORCID: https://orcid.org/0000-0002-2986 indicating the heterogeneous nature of DCM strength (Chen et al.
-9418 2016). Some empirical relations were proposed to predict the UCS
5
Postdoctoral Research Associate, Dept. of Civil and Environmental of cement–soil mixtures (Horpibulsuk et al. 2003; Lorenzo and
Engineering, Hong Kong Univ. of Science and Technology, Kowloon, Bergado 2004; Lee et al. 2005).
Hong Kong 999077, China. ORCID: https://orcid.org/0000-0002-0526
Although the relationships between DCM strength and strength-
-7471
Note. This manuscript was submitted on August 22, 2021; approved on improvement factors have been investigated, most were based on
April 25, 2022; published online on June 7, 2022. Discussion period open laboratory tests under controlled conditions, which could signifi-
until November 7, 2022; separate discussions must be submitted for indi- cantly deviate from conditions in the field. It is necessary to in-
vidual papers. This paper is part of the Journal of Geotechnical and vestigate and quantify the impacts of actual field conditions,
Geoenvironmental Engineering, © ASCE, ISSN 1090-0241. facilities, specifications, and techniques on DCM-treated offshore

© ASCE 04022063-1 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2022, 148(8): 04022063


soils. The significance level of impacts from the actual construc-
tion process may be evaluated based on the recorded construc-
tion data.
This study aimed to identify the mechanisms and key construc-
tion factors that influence DCM quality indicated by the homo-
geneity and magnitude of UCS. The construction process of a
practical offshore DCM project was investigated in close associa-
tion with cement–soil chemical and physical reactions. The main
objective was to assess the impact of key construction factors on the
DCM strengths in dominant types of natural marine deposits refer-
Downloaded from ascelibrary.org by Hong Kong University of Sci and Tech (HKUST) on 11/07/22. Copyright ASCE. For personal use only; all rights reserved.

ring to site investigation records, construction monitoring records


and quality assessment test results.
(a)

500
BH-5
DCM Construction
Test DCM
BH-4 Borehole
400
Longitude distance (m)

BH-3 Site Condition


300 The study site was at the Pearl River estuary in the south China
BH-2 BH-6 BH-7 Sea, where the stratigraphy of seabed soil can be divided into five
200
BH-1 categories along depth: marine clay, alluvial crust, lower allu-
vium deposit, completely decomposed granite, and bedrock.
100
BH-8
Subsoil profiles from eight boreholes at the site prior to DCM
construction are shown in Fig. 1(a). Soft silty clay and sandy–
0
0 200 400 600 800 1000 1200 silty clay dominated above −25 mPD, beneath which lay granular
Latitude distance (m) deposits, including sand, silt, and gravel. To lower disturbances to
(b) the marine ecological system, the seabed mud was not dredged
but improved with more than 4,000 DCM columns involving
Fig. 1. Study site information: (a) soil stratification records; and 630,000 m3 of marine soil. The properties of the marine deposits
(b) plan view of test DCM clusters and boreholes. prior to DCM construction are summarized in Table 1. The deposits
were mainly composed of clayey and silty soils with natural mois-
ture content occasionally over 100%. The average depth of DCM
clusters was −30mPD, with the majority located within the clay
Table 1. Properties of the marine deposit at study site layer. Selected DCM clusters were cored and tested after construc-
Properties Range tion for quality check. The layout of the test columns is shown in
Fig. 1(b).
Plastic limit (%) 21–59
Liquid limit (%) 44–91
Plasticity index (%) 22–59 DCM Construction Procedure
Clay content 0.39–0.55
Silt content 0.40–0.42 The offshore DCM was conducted using special barges equipped
Sand content 0.12–0.19 with mixing units, leaders, pumping units, slurry agitators, binder
Gravel content 0.01–0.09 silos, power plants, and operation rooms, shown in Fig. 2(a). Wet
Void ratio at p00 ¼ 1 kPa 1.54–1.93 cement slurry instead of dry powder was applied as the DCM
Compression index 0.42–0.53 cementing material in this project to achieve higher homogeneity
Natural moisture content (%) 43–124
of the stabilized soil, more uniform strength along depth, and
Saturated density (kN=m3 ) 12.5–18.4
smoother penetration in competent strata. The primary components

Fig. 2. Sketches of (a) DCM barge; (b) DCM mixing unit; and (c) arrangement of mixing shafts for a DCM cluster.

© ASCE 04022063-2 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2022, 148(8): 04022063


Table 2. Numbers of coring samples for quality testing
Curing age group
Dominant soil type 28–40 days 40–60 days 60–80 days
Clayey soils 1,071 623 446
Silty soils 48 28 38
Sandy soils 57 39 17
Gravelly soils 0 32 16
Total 1,176 722 517
Downloaded from ascelibrary.org by Hong Kong University of Sci and Tech (HKUST) on 11/07/22. Copyright ASCE. For personal use only; all rights reserved.

a single cluster in this study. The penetration was terminated when


Fig. 3. Pattern of installed DCM panels.
the mixing shafts reached the designed embedded depth within the
competent stratum. The bottom treatment was then conducted to
assure a sufficient shaft resistance of the DCM cluster in the com-
of the mixing unit are shown in Fig. 2(b). A single mixing cluster petent soil layer. Finally, in the withdrawal stage, the mixing unit
involved four shafts with an overlapped area between two adjacent mixed the cement slurry and the surrounding soil horizontally at a
clusters, as illustrated in Fig. 2(c). fixed rotating speed, providing a cross-section of the DCM clusters,
The DCM clusters were designed as separate panels to enhance as shown in Fig. 2(c). The cement slurry injection rate for bottom
the stability of the seawall in the tangent direction (Fig. 3). The plan treatment and withdrawal was set as equal, from 0.48 to 0.57 m3
view of the DCM panels and their designed embedded depths, lev- per minute. The penetrating and withdrawal mixing rates were es-
els of installation and competent stratum, required unconfined com- timated at 0.45 and 0.28 m per minute.
pressive strengths (RUCSs), and cement slurry injection rates are
shown in Fig. 4. The design installation depth followed the tentative
DCM Quality Tests
level of competent strata. The cement was ordinary portland cement.
In the mix design, the water–cement ratio (w=c) of the cement slurry After the completion of DCM construction, the stabilized cement–
was set at 0.8, which corresponded to a wet density of 1,641 kg=m3 . soil mixture was sampled and tested for uniformity, continuity,
The injected water during the penetration stage and natural soil water dimensions, and permeability. The verification of quality was con-
were not included in the design w=c ratio. The cement dosage for ducted with full-depth coring and unconfined compression tests at
mixing cement slurry was approximately 260 kg=m3. The volume various curing ages, from 28 to 80 days. The detailed sample sizes
fraction of injected cement slurry measures the cement content per are summarized in Table 2. The measured results and frequency
DCM cluster in practice. statistics are shown in Fig. 5. The Kolmogorov–Smirnov test was
The construction started by installing a geotextile layer and sand performed for normality checking, suggesting that the measure-
blanket. Subsequently, the DCM barge was anchored at a desig- ments do not strictly follow the normal distribution at confidence
nated location, followed by the penetration of the mixing units into level of 0.05. From the measured UCS values of cored DCM spec-
the marine soils with the mixing shafts rotating. During the pen- imens, the design strength requirements were exceeded substantially.
etration stage, the blades attached to the mixing shaft end broke The actual average UCS of 3.12 MPa was more than twice the design
and disturbed the soil, reducing its strength, allowing the mixing value [Fig. 4(d)]. The interval of the natural soil moisture content
shafts to be driven by their self-weight. The average volume frac- (Table 1) was also reduced from 0.43–1.24 to 0.036–0.956, demon-
tion of injected water during the penetration stage reached 12% for strating the effectiveness of DCM.

Fig. 4. Construction layout of DCM clusters in longitudinal direction: (a) design installation level; (b) design embedded depth; (c) tentative level of
competent stratum; (d) required unconfined compressive strength (RUCS); and (e) cement slurry injection rate.

© ASCE 04022063-3 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2022, 148(8): 04022063


0.16
Mean: 3.12 MPa
0.14 COV: 0.45

Relative Frequency
0.12 p-value (KS test): 1.31E-13

0.10
0.08
0.06
0.04
0.02
Downloaded from ascelibrary.org by Hong Kong University of Sci and Tech (HKUST) on 11/07/22. Copyright ASCE. For personal use only; all rights reserved.

0.00
0 2 4 6 8
(a)
UCS (MPa)

0.16
Mean: 45.95 %
0.14 COV: 0.27

Relative Frequency
0.12 p-value (KS test): 5.7E-6

0.10
0.08
0.06
0.04
0.02
0.00
0 20 40 60 80

(b) Moisture content (%)

0.16
Mean: 1639.43 kg/m3
0.14 COV: 0.046

Relative Frequency
0.12 p-value (KS test): 0.047

0.10
0.08
0.06
0.04
0.02
0.00
1400 1500 1600 1700 1800 1900 2000 2100

(c) Density (kg/m3)

Fig. 5. Quality assurance test results from sampled DCM clusters: (a) unconfined compressive stress (UCS); (b) moisture content; and (c) density.

Identification of Construction Impact Factors flocculation. Factors that impact the cation exchange with floccu-
lation during deep mixing include the natural soil type and water
Micromechanism of Cement–Soil Strength Formation content.
With the calcium ions continuously accumulating, the active cal-
Micromechanisms of strength formation of cement-stabilized soil cium ions can react with the silica and alumina from the soil in an
can refer to the chemical reactions (i.e. cement hydration, cation alkaline environment, generating hydrated calcium silicate (CSH)
exchange, and pozzolanic reactions) and physical contacts between and hydrated calcium aluminate (CAH) cementitious matrices with
soil particles and cement slurry shown in Fig. 6. Cement hydration a fibrous crystal structure (Sherwood 1993). The exposed positive
is a collection of chemical reactions operating in series, primarily ions continuously adsorb hydroxide ions, forming solidification
the reaction between water and the cement calcium silicate with wrapping around the soil particle. This process is the so-called poz-
tricalcium aluminate, which produce an alkaline environment with zolanic reaction, which is sensitive to soil type with different con-
calcium hydroxide, hydrated calcium silicate, and calcium alumi- centrations of silica, alumina, and water as reactants. Furthermore,
nate gels that harden with time (Bullard et al. 2011). The amounts the long-period reaction in an alkali environment can be impacted
of basic reactants, such as cement dosage, cement slurry quality, by disturbance to the alkalinity during DCM construction.
and water volume, can influence the hydration reactions. Among the previous chemical reactions, physical contact
Negatively charged silicate particles from natural marine soils [Fig. 6(b)] plays an important role. Various contact degrees of ce-
are attached with some cations, including Naþ and Kþ . Equivalent ment slurry with soil particles can occur during DCM construction.
exchange can occur immediately between these positive ions and When the soil particle surface was isolated from the adhesion of
calcium ions released from the calcium hydroxide, reducing water other substances, the voids between them were filled with cement
adsorption of the clay particles. The increased electric charge de- slurry, leading to an ideal reinforcement effect. The soil grains and
creases the repulsion between the negatively charged clay particles pore water can obstruct the pore channel, resulting in partial filling
and the thickness of the adsorbed water layer, gradually forming with less contact. Some soil aggregates may not be fully crushed

© ASCE 04022063-4 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2022, 148(8): 04022063


Downloaded from ascelibrary.org by Hong Kong University of Sci and Tech (HKUST) on 11/07/22. Copyright ASCE. For personal use only; all rights reserved.

Fig. 6. Mechanism of strength gain of cement stabilized soil referring to: (a) chemical reactions; and (b) physical contact between cement slurry and
soil aggregates.

during the penetration and mixing process, and the cement slurry higher clay fraction adsorb metal ions better than sandy soils, lead-
only wraps on their interfaces without filling the voids. ing to a higher concentration of adsorbed calcium ions during cat-
ion exchange (e.g., Manrique et al. 1991). Furthermore, the varying
silica and aluminate and pH levels in different soils can affect the
Construction Impact Factors pozzolanic reaction. The soil pore structure provides distinct per-
The UCS of DCM mixtures can be influenced by binder character- meability properties, which may also affect the heterogeneity of the
istics, soil conditions, mixing conditions, and curing conditions strength of DCM treated soil. Macropores (>0.075 mm) are more
(Terashi 1997) based on laboratory studies. However, many of the abundant in coarse-grained soils (sandy soils), accounting for about
conditions cannot be well controlled in field conditions, and how 35%–50% of the overall pore volume, whereas fine-grained soils
the strength of cement-stabilized soil behavior is influenced by con- (silty and clayey soils) are dominated by micropores (<0.03 mm)
struction techniques and site conditions has not been fully explored and have smaller particle spacing (SSGTC and SSSA 2008).
in the previously reported studies in the literature. In this study, The key factors related to the construction procedure include the
given the illustrated DCM construction workflow and the mecha- volume fraction of injected water, volume fraction of injected
nisms of DCM strength formation shown in Fig. 7, several key con- cement slurry, cement injection rate (m3 =min), and penetrating and
struction factors that may influence the DCM final strength are mixing time per meter (min=m). The volume fraction of injected
identified. These factors are associated with construction site con- water is the ratio of the injected water volume during penetration
ditions, construction procedures, and curing conditions. to the bulk volume of a single DCM cluster, representing the water
Impact factors related to construction site conditions include quantity for each cluster. Water is a reactant involved in hydration,
fluctuation of tidal level (m/hour) and soil type. Large fluctuations cation exchange, and pozzolanic reactions. Nevertheless, excessive
of tidal level can disturb the mixing shafts and blades during the water can reduce the concentration of other reactants and the envi-
penetration, lifting, and mixing stages of DCM construction, lead- ronmental alkalinity in chemical reactions and increase water pock-
ing to partial contact between cement slurry and soil aggregates. ets not filled by cement slurry, leading to strength reduction. The
Different soil types provide varied cation exchange capacities volume fraction of injected cement slurry indicates the cement
because the cations can be held by the negatively charged particles quantity as a basic reactant for hydration. The produced calcium
of fine-grained soils with thin and plate structures, and sand cannot hydroxide also participates in subsequent cation exchange and
exchange cations without electrical charge. Hence, soils with a pozzolanic reactions, contributing to the strength gain of the

© ASCE 04022063-5 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2022, 148(8): 04022063


Downloaded from ascelibrary.org by Hong Kong University of Sci and Tech (HKUST) on 11/07/22. Copyright ASCE. For personal use only; all rights reserved.

Fig. 7. Workflow of DCM mixing unit and associated key factors affecting strength formation.

cement–soil mixture. A higher cement injection rate can provide 28 to over 80 days in 20-day intervals. The increase of cement-
greater initial kinetic energy of cement slurry from nozzles, facili- stabilized soil strength tended to level off beyond 60 days when
tating the cement slurry to permeate into the voids between soil the initial water content of the original soil was high. The whole
particles more thoroughly at the bottom treatment, lifting, and mix- DCM construction was performed in the marine area and under the
ing stages. At the penetrating and mixing stages, the mixing units seabed level, and water was injected while penetrating the soil for
crushed the natural soil mass into small aggregates through cutting protecting the mixing units and smoothing the penetration process.
blades, allowing sufficient contact between the soil particles and the The injected water and cement slurry volumes were recorded in the
injected cement slurry. Increasing penetration and mixing time can penetrating and mixing stages. These volumes were transformed
increase the contact surface area and accelerate the second phase of into volume fractions by dividing the individual bulk volume of
the cement hydration and pozzolanic reaction. the tested DCM column to normalize the influence of the DCM
The influence of the curing condition is observed through test cluster size.
samples’ age and moisture content. The pozzolanic reaction can The scatterplots of the average UCS versus seven influence fac-
last months to years for long-term strength formation, leading to tors for marine clay are presented in Fig. 8, reflecting the quantity
the variation of sample strength at different ages. The moisture and fluctuation of UCS with these factors. The coefficients of varia-
contents of cored samples indicate wetness at the curing stage. tion (COV, ratio of standard deviation and mean) of UCS associated
Excessive water can reduce the alkalinity and concentration of with these factors are shown in Fig. 9. The COV reflects the homo-
silica and alumina for pozzolanic reactions during the curing stage. geneity level of UCS in the construction site. A greater value of
Considering water injection during construction, the multiple COV indicates decreased homogeneity of UCS and greater uncer-
stages of pozzolanic reaction, and the hydrate crystal products, tainty from the influence factors. The average UCS and the COV
the curing moisture content at time of coring is used as a direct for dominant natural soil types are summarized in Fig. 10.
field indicator to indicate the water content in DCM samples rather
than the moisture content in natural soil. Importance of Influence Factors
Sensitivity analysis was performed to identify the key construction
Interpretation of Construction Effects Based on factors that impact the strength and uniformity of DCM construc-
Field Records tions, based on the out-of-bag (OOB) importance estimation. The
OOB estimation measures how influential the predictor variables
In order to quantify the influence of each key in-situ construction (i.e., DCM construction factors) are in predicting the response
factor, project field records were analyzed along with the construc- (i.e. observed UCS) by permutation using a random forest model
tion conditions and measured DCM strengths. (e.g., Altmann et al. 2010). The influence of a particular construc-
tion factor can be manifested by the variation of model error by
random permutations. If a construction factor has a greater influ-
Processing of Field Records
ence on the UCS, then permutating its value should affect the
According to the site borehole records and geotechnical site inves- model error significantly, whereas if a construction factor is not
tigations, the average UCS of cored DCM-stabilized soil mixtures influential, there will be little variation of the model error after per-
was calculated for four principal soil types: clayey, silty, sandy, and mutating its value. The error difference assuming a random forest
gravelly soils. This study grouped curing ages into four sets from model of T trees and P predictors is shown as follows:

© ASCE 04022063-6 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2022, 148(8): 04022063


Pn Pn observations for each predictor variable j ¼ 1; : : : ; P, the gener-
ðyi − ŷbag;tj Þ2 ðyi − ŷtj Þ2
dtj ¼ εtj − εt ¼ i¼1
− i¼1
ð1Þ ated UCS ŷbag and model error εtj can be estimated. The OOB pre-
n n
dictor importance can be quantified by the ratio between the
where for each tree t ¼ 1; : : : ; T; n = number of samples; εt = OOB difference of errors averaged over all trees and the standard
error estimated by the mean squared error; yi = observed UCS; and deviation dj =σj . The larger the ratio dj =σj , the more influence
ŷ = predicted UCS from the model. After randomly permutating the the key factor has on the DCM. The results of the predictor

8 8 8 8
Downloaded from ascelibrary.org by Hong Kong University of Sci and Tech (HKUST) on 11/07/22. Copyright ASCE. For personal use only; all rights reserved.

7 7 7 7
Average UCS (MPa)

6 6 6 6
5 5 5 5
4 4 4 4
3 3 3 3
2 2 2 2
1 1 1 1
0 0 0 0
0.2 0.3 0.4 0.48 0.50 0.52 0.54 0.56 0.58 0.05 0.10 0.15 0.20 0.25 0.30 0.27 0.30 0.33 0.36 0.39 0.42
Tidal fluctuation (m/h) Injection rate (m3 /min) Injected water volume fraction Injected slurry volume fraction
(a) (b) (c) (d)
8 8 8
7 7 7
Average UCS (MPa)

6 6 6
5 5 5
4 4 4
3 3 3
2 2 2
1 1 1
0 0 0
1.0 1.5 2.0 2.5 3.0 3.5 4.0 3.0 3.3 3.6 3.9 10 20 30 40 50 60 70 80
Penetration unit time (min/m) Mixing unit time (min/m) Moisture content (%)
(e) (f) (g)

Fig. 8. Relationships between the average UCS of marine clay and key DCM construction factors: (a) tidal fluctuation (m=h); (b) injection rate
(m3 =min); (c) injected water volume fraction; (d) injected slurry volume fraction; (e) penetration unit time (min/m); (f) mixing unit time (min/m); and
(g) moisture content (%).

1.0 1.0 1.0 1.0

0.8 0.8 0.8 0.8


COV of UCS

0.6 0.6 0.6 0.6

0.4 0.4 0.4 0.4

0.2 0.2 0.2 0.2

0.0 0.0 0.0 0.0


0.2 0.3 0.4 0.48 0.50 0.52 0.54 0.56 0.58 0.05 0.10 0.15 0.20 0.25 0.30 0.27 0.30 0.33 0.36 0.39 0.42
Tidal fluctuation (m/h) Injection rate (m3 /min) Injected water volume fraction Injected slurry volume fraction
(a) (b) (c) (d)
1.0 1.0 1.0

0.8 0.8 0.8


COV of UCS

0.6 0.6 0.6

0.4 0.4 0.4

0.2 0.2 0.2

0.0 0.0 0.0


1.0 1.5 2.0 2.5 3.0 3.5 4.0 3.0 3.3 3.6 3.9 10 20 30 40 50 60 70 80
Penetration unit time (min/m) Mixing unit time (min/m) Moisture content (%)
(e) (f) (g)

Fig. 9. Relationships between the COV of UCS of marine clay and key DCM constructions factors: (a) tidal fluctuation (m=h); (b) injection rate
(m3 =min); (c) injected water volume fraction; (d) injected slurry volume fraction; (e) penetration unit time (min/m); (f) mixing unit time (min/m); and
(g) moisture content (%).

© ASCE 04022063-7 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2022, 148(8): 04022063


4 28-40days 0.7 28-40days
40-60days 40-60days
60-80days 0.6 60-80days

Average UCS (MPa)


3
0.5

COV of UCS
0.4
2
0.3

0.2
1
Downloaded from ascelibrary.org by Hong Kong University of Sci and Tech (HKUST) on 11/07/22. Copyright ASCE. For personal use only; all rights reserved.

0.1

0 0.0
Clayey
CLAY Silty
SILT Sandy
SAND Gravelly
GRAVEL Clayey
CLAY Silty
SILT Sandy
SAND Gravelly
GRAVEL
soils soils soils soils soils soils soils soils
Dominant soil type Dominant soil type

Fig. 10. Average and COV of UCS from DCM samples for varied natural dominant soil types.

Fig. 11. Out-of-bag permuted predictor importance estimates for key factors from offshore DCM construction. A larger value suggests a greater
influence.

importance estimates for all key factors are summarized in Fig. 11. cation exchange, which further influences the pozzolanic reactions
As can be seen, for the DCM offshore construction project, the that form long-term strength. There was an ascending trend in the
amount of water injected during drilling and type of native soil COV from marine clay to gravel, indicating that the homogeneity
are the most important factors, whereas the difference in grouting level of UCS becomes lower with increasing particle size or coarse
rate in the range of 0.48–0.58 m3 =min does not affect the strength content. It is reasonable because large voids are more likely to form
substantially. when the soil aggregates are large, which easily introduce varied
cement slurry saturation degrees when the same construction tech-
Influence of Key Factors nique is applied to all soil types during DCM construction. An ex-
ception was the COV for silt soil, which was larger than the
The influence of nine key factors on the UCS was demonstrated
COV for sandy soil. One explanation is that silt soil has a dual-
referring to the strengths of cored samples and strength gain mech-
porosity structure compared to clays and sand, and its mechanical
anisms of cement–soil mixtures.
Original dominant soil type. Based on borehole records, four behavior is controlled by soil aggregates and coarse particles
dominant soil types were involved: clayey, silty, sandy, and grav- together (e.g., Zhao et al. 2013). This structure with large interag-
elly soils. The average UCS of the DCM samples was approxi- gregate pores leads to varying degrees of permeation of cement
mately 2.5 MPa. The UCS of the samples originating in marine slurry into the voids during construction [Fig. 6(b)], possibly lead-
clay was the highest, leading by almost 0.5 MPa for all ages com- ing to heterogeneity in the treated soil.
pared to DCM mixtures in other soil types [Fig. 10(a)]. The inher- Fluctuation of tidal level. Another potential disturbance of off-
ent difference in dominant soils’ cation exchange capacity (CEC) shore DCM construction is the tidal impact on the mixing units on
could be considered a contributing factor for strength gains in dif- the DCM barge. Reduction of the average UCS and large fluctua-
ferent types of soil. The marine clays had the highest clay fraction, tions in the COV could be observed after tidal fluctuation reached
composed of negatively charged clay particles with greater CEC 0.38 m per hour [Fig. 8(a)]. Large tidal fluctuations may introduce
than other dominant soil types. Therefore, a higher concentration vibrations of the mixing shafts and blades that operate in the lifting
of calcium ions adsorbed to clay particles can be achieved from and rotating stages. Such vibrations may disturb the mixing unit

© ASCE 04022063-8 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2022, 148(8): 04022063


to varying extents and produce uneven contact conditions between associated with the strength gain of cement–soil mixtures. The in-
the soil particles and cement slurry, leading to uncertainty in the creased mixing time also reduces the incomplete filling of the voids
homogeneity of cement-stabilized soil. between soil aggregates by the cement slurry, leading to improved
Volume fraction of injected water. The volume fraction of in- homogeneity.
jected water, primarily in the range of 0.05–0.30, represents the Age. Cement-stabilized soil strength grows slowly over 60 days
volume of injected water for protecting the mixing unit and when the initial water content of the natural soils is high (Kawasaki
smoothening the penetration activity. It does not include the water et al. 1981). Indeed, in this project, the magnitude and homogeneity
in the natural soils and cement slurry. The average UCS decreases level of strength increased substantially during the curing ages of
with increasing injected water volume, from an initial 3.5 MPa to 28–60 days for marine clayey, silty, and sandy soils, as shown in
around 2.5 MPa [Fig. 8(c)]. Although the injected water constitutes Fig. 10. With the increasing curing age, the pozzolanic reactions
Downloaded from ascelibrary.org by Hong Kong University of Sci and Tech (HKUST) on 11/07/22. Copyright ASCE. For personal use only; all rights reserved.

the primary reactant for the hydration and pozzolanic reactions in tend to complete, with more cemented calcium silicate and calcium
the cement soil mixtures, excessive water injection can reduce the aluminate hydrate accumulating and surrounding the soil particles,
alkalinity of the environment and concentration of other reactants improving the cementation strength. After 60 days, the UCS for
involved in the cation exchange, hydration, and pozzolanic reac- the silty and gravelly marine soils reached a stabilized level, but
tions. The redundant water contributes to the separation of soil those for the clayey and sandy soils continued to grow. This can be
aggregates by water pockets, leading to reduced UCS. explained by the difference in the pore structure. Both the dual-
Volume fraction of injected cement slurry. The average UCS porosity structure of the silty soil with coarse particles and the
increased with the injected volume fraction of cement slurry in the macropores of gravelly soils introduce high permeability character-
range of 0.27–0.42 [Fig. 8(d)]. Cement slurry is the main reactant istics, allowing ease of fluid flow. With water infiltration, the envi-
for generating hydrated calcium-based gels and forming cementi- ronmental alkalinity decreases, and the concentration of activated
tious matrices with a fibrous crystal structure, which dominates the calcium ions is diluted. Therefore, the reduced concentration of
strength formation of cemented soils. The increased volume frac- reactants in the pozzolanic reaction leads to a mitigated reaction
tion of cement slurry directly improves the principal components degree and stabilized strength.
contributing to the strength. The COV reached a minimum value at Moisture content. The moisture content of the cement–soil
around 0.2 when the injected slurry increased to 0.41 [Fig. 9(d)]. mixture was measured during the UCS test [Figs. 8(g) and 9(g)].
Injection of sufficient cement slurry ensures the cement slurry fills The average UCS gradually decreased to around 2 MPa at a
the off-contact gaps between soil aggregates, contributing to higher moisture content of 80% in the cement–soil mixtures. The initial
physical contact degree and homogeneity of strength. The COV water in the cement–soil mixture serves as a reactant for pozzolanic
was high initially. When the amount of cement slurry injected is reactions. However, the excessive amount of water can decrease the
low, only part of the pores in the soil are filled, thus producing concentration of reactants and the alkaline chemical reaction envi-
a nonuniform distribution of hydration products. When the injected ronment, which then delays the long-term strength formation pro-
volume rises, more pores are filled and the reactants are in good cess of cement-stabilized soils. Large fluctuations in COV are
contact, leading to a reduced COV. observed after the moisture content is greater than 60% for different
Cement slurry injection rate. A higher cement slurry injection curing ages. High moisture content can cause uneven distribution
rate provides larger kinetic energy of cement slurry from the noz- of excessive water, which requires additional time to react with the
zles of the mixing units, improving the contact degree of cement silica and alumina from the soil and calcium hydroxide during
slurry and soil aggregates. The injection rate ranges between 0.48 pozzolanic reactions, leading to the varying COV at different curing
and 0.57 m3 =min, leading to approximately 0.35 m3 =m of the ages.
grout volume fraction under identical mixing time per meter. The
average UCS and COV [Figs. 8(b) and 9(b)] varied little within this
range of injection rate, as did the grout penetration in soil aggre- Summary and Conclusions
gates. The effect of the change in grout volume per unit time on the
UCS can be obscured by the variation of the volume fraction of Compared to traditional dredged land reclamation methods, off-
cement slurry. Hence this factor is the least sensitive construction shore DCM construction is more environmentally friendly without
factor in terms of the estimated OOB importance (Fig. 11). dredging or disturbance to the marine ecological system. An off-
Penetrating time per meter. When mixing units are drilled shore DCM project for undredged land reclamation was evaluated
through a stiff soil layer or a hard obstruction during the penetration to identify key construction factors that affect the homogeneity and
stage, additional time is required, and a larger amount of water will magnitude of UCS. These factors are related to geological condi-
be injected to protect the mixing units and smooth the penetration tions, construction procedures, curing conditions, and chemical and
process. Therefore, the average UCS and COV variations with pen- physical mechanisms of cement-stabilized soils. Several conclu-
etration time per meter [Figs. 8(e) and 9(e)] are similar to those of sions can be drawn:
the injected water volume fraction. As mentioned, the drilling pro- 1. The impact on the UCS from the volume fraction of injected
cess broke up large soil clods into smaller pieces and improved the water and the natural moisture content is most significant among
contact conditions of the cementation reactants. However, the drill- all the evaluated factors from offshore DCM construction. The
ing process was accompanied by water injection, so its effect on volume of injected water must be carefully calibrated, especially
strength was overshadowed by the dominant factor, water volume, when the marine soils to be treated have high natural water con-
as illustrated in Fig. 11. tents. Additional injected water at the penetration stage may ex-
Mixing time per meter. With the increased mixing time from ceed the hydration and pozzolanic requirements and decrease
3.1 to 3.9 min per meter, the average UCS improved from an initial the concentration of other reactants and alkalinity of the envi-
3 to 3.5 MPa [Fig. 8(f)], and the homogeneity level of UCS also ronment, affecting the strength formation of cement-stabilized
improved, indicated by a decrease in the COV [Fig. 9(f)]. The soils.
increased mixing time per meter ensures the mixing blades thor- 2. The varying soil types at the construction site significantly im-
oughly cut the original soil aggregates into small pieces and pacted the UCS of offshore DCM in terms of both homogeneity
increases the contact area of the reactants for chemical reactions and magnitude. Due to the inherent variation of the cation

© ASCE 04022063-9 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2022, 148(8): 04022063


exchange capability, fine-grained marine soils with high clay improved ground.” Comput. Geotech. 43 (Jun): 37–50. https://doi
fractions generate greater cement–soil mixture strength com- .org/10.1016/j.compgeo.2012.02.003.
pared to coarse-grained sandy soils. Hence specific injection Horpibulsuk, S., N. Miura, and T. S. Nagaraj. 2003. “Assessment of
and mixing methodologies should be recommended for different strength development in cement-admixed high-water content clays with
Abrams’ law as a basis.” Géotechnique 53 (4): 439–444. https://doi.org
soils, including using extra mixing time and cement slurry in-
/10.1680/geot.2003.53.4.439.
jection to ensure adequate cutting effects of mixing blades and
Jiang, Y., J. Han, and G. Zheng. 2013. “Numerical analysis of consolidation
increase the contact degree between the cement slurry and soil of soft soils fully penetrated by deep-mixed columns.” KSCE J. Civ.
aggregates. Eng. 17 (1): 96–105. https://doi.org/10.1007/s12205-013-1641-x.
3. Longer mixing time during construction did not reduce the Kawasaki, T., A. Niina, S. Saitoh, Y. Suzuki, and Y. Honjo. 1981. “Deep
UCS but promoted it, whereas large tidal-level fluctuations mixing method using cement hardening agent.” In Vol. 3 of Proc., 10th
Downloaded from ascelibrary.org by Hong Kong University of Sci and Tech (HKUST) on 11/07/22. Copyright ASCE. For personal use only; all rights reserved.

may lead to vibrations of the mixing units and produce strength Int. Conf. on Soil Mechanics and Found. Eng., 721–724. London:
variations. ISSMGE.
4. The strength of DCM increases with curing age within a curing Kitazume, M., M. Grisolia, E. Leder, I. P. Marzano, A. A. S. Correia, P. J. V.
age up to 80 days for the dominant clayey and sandy marine Oliveira, and M. Andersson. 2015. “Applicability of molding proce-
soils. For silty and gravelly soils, the UCS appeared to reach dures in laboratory mix tests for quality control and assurance of the
deep mixing method.” Soils Found. 55 (4): 761–777. https://doi.org/10
a stabilized level 60 days after the DCM treatment.
.1016/j.sandf.2015.06.009.
Larsson, S., M. Dahlström, and B. Nilsson. 2005a. “A complementary field
study on the uniformity of lime-cement columns for deep mixing.”
Data Availability Statement Proc. Inst. Civ. Eng. Ground Improv. 9 (2): 67–77. https://doi.org/10
.1680/grim.2005.9.2.67.
Some or all data, models, or code generated or used during the Larsson, S., M. Dahlström, and B. Nilsson. 2005b. “Uniformity of lime-
study are available from the corresponding author by request. cement columns for deep mixing: A field study.” Proc. Inst. Civ. Eng.
Ground Improv. 9 (1): 1–15. https://doi.org/10.1680/grim.2005.9.1.1.
Lee, F. H., C. H. Lee, and G. R. Dasari. 2006. “Centrifuge modelling of wet
Acknowledgments deep mixing processes in soft clays.” Géotechnique 56 (10): 677–691.
https://doi.org/10.1680/geot.2006.56.10.677.
The work presented in this paper was substantially supported by Lee, F. H., Y. Lee, S. H. Chew, and K. Y. Yong. 2005. “Strength and modulus
Eunsung O&C Offshore Marine and Construction (No. EUNSUN- of marine clay–cement mixes.” J. Geotech. Geoenviron. Eng. 131 (2):
G19EG01). 178–186. https://doi.org/10.1061/(ASCE)1090-0241(2005)131:2(178).
Liu, S. Y., Y. J. Du, Y. L. Yi, and A. J. Puppala. 2012. “Field investigations
on performance of T-shaped deep mixed soil cement column–supported
embankments over soft ground.” J. Geotech. Geoenviron. Eng. 138 (6):
References 718–727. https://doi.org/10.1061/(ASCE)GT.1943-5606.0000625.
Abusharar, S. W., J. J. Zheng, and B. G. Chen. 2009. “Finite element mod- Lorenzo, G. A., and D. T. Bergado. 2004. “Fundamental parameters of
elling of the consolidation behavior of multi-column supported road cement-admixed clay—New approach.” J. Geotech. Geoenviron. Eng.
embankment.” Comput. Geotech. 36 (4): 676–685. https://doi.org/10 130 (10): 1042–1050. https://doi.org/10.1061/(ASCE)1090-0241(2004)
.1016/j.compgeo.2008.09.006. 130:10(1042).
Altmann, A., L. Toloşi, O. Sander, and T. Lengauer. 2010. “Permutation Manrique, L. A., C. A. Jones, and P. T. Dyke. 1991. “Predicting cation-
importance: A corrected feature importance measure.” Bioinformatics exchange capacity from soil physical and chemical properties.” Soil
26 (10): 1340–1347. https://doi.org/10.1093/bioinformatics/btq134. Sci. Soc. Am. J. 55 (3): 787–794. https://doi.org/10.2136/sssaj1991
Bellato, D., I. P. Marzano, and P. Simonini. 2020. “Microstructural analyses .03615995005500030026x.
of a stabilized sand by a deep-mixing method.” J. Geotech. Geoenviron. Mohammadinia, A., M. M. Disfani, D. Conomy, A. Arulrajah, S. Horpibulsuk,
Eng. 146 (6): 04020032. https://doi.org/10.1061/(ASCE)GT.1943-5606 and S. Darmawan. 2019. “Utilization of alkali-activated fly ash for
.0002254. construction of deep mixed columns in loose sands.” J. Mater. Civ.
Bullard, J. W., H. M. Jennings, R. A. Livingston, A. Nonat, G. W. Scherer, Eng. 31 (10): 04019233. https://doi.org/10.1061/(ASCE)MT.1943-5533
J. S. Schweitzer, and J. J. Thomas. 2011. “Mechanisms of cement .0002878.
hydration.” Cem. Concr. Res. 41 (12): 1208–1223. https://doi.org/10 Navin, M. P. 2005. “Stability of embankments founded on soft soil
.1016/j.cemconres.2010.09.011. improved with deep-mixing-method columns.” Doctoral Dissertation,
Chen, E. J., Y. Liu, and F. H. Lee. 2016. “A statistical model for the Dept. of Civil Engineering and Environmental Engineering, Virginia
unconfined compressive strength of deep-mixed columns.” Géotechnique Tech.
66 (5): 351–365. https://doi.org/10.1680/jgeot.14.P.162. Rogers, C. D. F., and S. Glendinning. 1997. “Improvement of clay soils in
Chen, J., F. H. Lee, and C. C. Ng. 2011. “Statistical analysis for strength situ using lime piles in the UK.” Eng. Geol. 47 (3): 243–257. https://doi
variation of deep mixing columns in Singapore.” In Geo-frontiers: .org/10.1016/S0013-7952(97)00022-7.
Advances in geotechnical engineering, GSP No. 211, edited by J. Han Sherwood, P. T. 1993. Soil stabilization with cement and lime—State of the
and D. E. Alzamora, 576–584. Reston, VA: ASCE. art review. London: Transport Research Laboratory, Dept. of Transport,
Disfani, M. M., A. Mohammadinia, A. Arulrajah, S. Horpibulsuk, and HMSO publications.
M. Leong. 2021. “Lightly stabilized loose sands with alkali-activated Soil Science Glossary Terms Committee and Soil Science Society of
fly ash in deep mixing applications.” Int. J. Geomech. 21 (3): 04021011. America. 2008. Glossary of soil science terms 2008. Madison, WI: Soil
https://doi.org/10.1061/(ASCE)GM.1943-5622.0001958. Science Society of America.
Furudoi, T. 2005. “Second phase construction project of Kansai International Subramaniam, P., and S. Banerjee. 2020. “Dynamic properties of cement-
Airport—Large-scale reclamation works on soft deposits.” In Vol. 16 of treated marine clay.” Int. J. Geomech. 20 (6): 04020065. https://doi.org
Proc. of the Int. Conf. on Soil Mechanics and Geotechnical Engineering, /10.1061/(ASCE)GM.1943-5622.0001673.
313. Amsterdam, Netherlands: IOS Press BV. Terashi, M. 1997. “Theme lecture: Deep mixing method—Brief state of the
Honjo, Y. 1982. “A probabilistic approach to evaluate shear strength of art.” In Vol. 4 of Proc. 14th Int. Conf. on Soil Mech. and Found. Eng.,
heterogeneous stabilized ground by deep mixing method.” Soils Found. 2475–2478. Rotterdam, Netherlands: A.A. Balkema.
22 (1): 23–38. https://doi.org/10.3208/sandf1972.22.23. Watabe, Y., and T. Noguchi. 2011. “Site-investigation and geotechnical de-
Horpibulsuk, S., A. Chinkulkijniwat, A. Cholphatsorn, J. Suebsuk, and sign of D-runway construction in Tokyo Haneda airport.” Soils Found.
M. D. Liu. 2012. “Consolidation behavior of soil-cement column 51 (6): 1003–1018. https://doi.org/10.3208/sandf.51.1003.

© ASCE 04022063-10 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2022, 148(8): 04022063


Wijerathna, M., and D. S. Liyanapathirana. 2018. “Reliability-based per- Yapage, N. N. S., D. S. Liyanapathirana, R. B. Kelly, H. G. Poulos, and
formance of embankments improved with deep mixing considering spa- C. J. Leo. 2014. “Numerical modeling of an embankment over soft
tial variability of material properties.” ASCE-ASME J. Risk Uncertainty ground improved with deep cement mixed columns: Case history.”
Eng. Syst. Part A: Civ. Eng. 4 (4): 04018035. https://doi.org/10.1061 J. Geotech. Geoenviron. Eng. 140 (11): 04014062. https://doi.org/10
/AJRUA6.0000987. .1061/(ASCE)GT.1943-5606.0001165.
Wijerathna, M., and D. S. Liyanapathirana. 2019. “Significance of spatial Yin, J. H., and Z. Fang. 2006. “Physical modelling of consolidation behav-
variability of deep cement mixed columns on reliability of column- ior of a composite foundation consisting of cement-mixed soil column
supported embankments.” Int. J. Geomech. 19 (8): 04019087. https://doi and untreated soft marine clay.” Geotechnique 56 (1): 63–68. https://doi
.org/10.1061/(ASCE)GM.1943-5622.0001473. .org/10.1680/geot.2006.56.1.63.
Yang, T., J. Z. Yang, and J. Ni. 2014. “Analytical solution for the consoli- Zhao, H. F., L. M. Zhang, and D. S. Chang. 2013. “Behavior of coarse
dation of a composite ground reinforced by partially penetrated imper- widely graded soils under low confining pressures.” J. Geotech. Geo-
Downloaded from ascelibrary.org by Hong Kong University of Sci and Tech (HKUST) on 11/07/22. Copyright ASCE. For personal use only; all rights reserved.

vious columns.” Comput. Geotech. 57 (Apr): 30–36. https://doi.org/10 environ. Eng. 139 (1): 35–48. https://doi.org/10.1061/(ASCE)GT.1943
.1016/j.compgeo.2014.01.001. -5606.0000755.

© ASCE 04022063-11 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2022, 148(8): 04022063

You might also like