Sie sind auf Seite 1von 7

Alkali-Activated Ground-Granulated Blast Furnace Slag for

Stabilization of Marine Soft Clay


Yaolin Yi 1; Cheng Li 2; and Songyu Liu, M.ASCE 3
Downloaded from ascelibrary.org by Sardar Vallabhbhai Natl Inst Of Tech (svnit) on 07/27/16. Copyright ASCE. For personal use only; all rights reserved.

Abstract: This paper investigated the stabilization efficacy of alkali-activated ground-granulated blast furnace slag (GGBS) for a marine soft
clay, compared with that of portland cement (PC). The influence of activators, including NaOH, Na2 CO3 , carbide slag (CS), NaOH-CS,
Na2 CO3 -CS, and Na2 SO4 -CS, on the stabilization efficacy was investigated. A range of tests were conducted to investigate the properties of
stabilized clays, including unconfined compressive strength (UCS), X-ray diffraction (XRD), scanning electron microscopy (SEM), and
mercury intrusion porosimetry (MIP). The results indicated that Na2 CO3 -GGBS had no stabilization efficacy for this marine soft clay.
NaOH-GGBS-stabilized clay yielded the highest UCS at 7, 28, and 90 days; however, the UCS decreased from 90 to 180 days because
of the microcracking. CS-GGBS-stabilized clay had higher 90-day and 180-day UCS than that of PC-stabilized clay, but significantly lower
7-day and 28-day UCS. NaOH, Na2 CO3 , and Na2 SO4 could enhance the strength development rate of CS-GGBS-stabilized clay. However,
the UCS of NaOH-CS-GGBS and Na2 CO3 -CS-GGBS-stabilized clays decreased from 90 to 180 days as well. Na2 SO4 -CS-GGBS was found
to be the optimum binder for this marine soft clay, yielding at least twice higher UCS than that of PC stabilized clay at any age studied.
Considerable ettringite was produced in the Na2 SO4 -CS-GGBS stabilized clay, which contributed to the enhanced strength. DOI: 10.1061/
(ASCE)MT.1943-5533.0001100. © 2014 American Society of Civil Engineers.
Author keywords: Soft clay; Alkali activation; Ground-granulated blast furnace; Unconfined compressive strength; Microstructure.

Introduction furnace slag, (GGBS) has gained popularity (Roy 1999; Shi
et al. 2006).
There is a large amount of marine soft clay, with high water con- GGBS is a by-product of the steel industry, which consists of a
tent, high compressibility, and low shear strength, along the coast substantial proportion of a glassy phase with a substantial content
of China. The deep cement mixing method is one of the most of Ca, Si, Al, and Mg-based compounds, most likely being present
widely used marine soft clay improvement methods in China; it as a solid solution, with potential cementitious reactivity. The
mixes portland cement (PC) with in situ soft clay to form stabilized generation of 1 t of GGBS consumes only approximately 1,300 MJ
clay with high strength (Bergado et al. 1996; Porbaha 1998; Holm energy and generates 0.07 t of CO2 emissions (Higgins 2007).
2003; Terashi 2003; Kitazume and Terash 2013). The PC clay sta- Currently, the price of GGBS in China is 60–80% that of PC.
bilization is primarily through the PC hydration reactions, leading GGBS needs to be chemically activated to increase its hydration
to the formation of calcium silicate hydrates (CSH), calcium alu- rate and produce compounds such as CSH, CAH, or CASH (Wang
minates hydrates (CAH), and calcium aluminum silicate hydrates and Scrivener 1995; Song et al. 2000; Shi et al. 2006). A range of
(CASH) (Bergado et al. 1996; Porbaha et al. 2000; Chew et al. chemicals can be used as GGBS activators, among which NaOH,
2004; Kitazume and Terash 2013). However, there are significant Na2 CO3 , Na2 O · nSiO2 (water glass), and Na2 SO4 are the most
environmental effects associated with PC production in terms of widely available and economical (Shi et al. 2006). Previously
high CO2 emissions (approximately 0.95 t CO2 =t PC), energy reported results from Russia and China have suggested that the best
consumption (approximately 5,000 MJ=t PC), and consumption activator for compressive strength was water glass solution
of nonrenewable resources (approximately 1.5 t limestone and (Bakharev et al. 1999); however, water glass usually results in fast
clay/t PC) (Higgins 2007). Portland cement production is setting and thereby causes mixing and placement problems (Shi
responsible for 5–8% of artificial CO2 emissions (Scrivener and et al. 2006). The activation mechanism of the GGBS is very com-
Kirkpatrick 2008). Hence, the development of alternative sustain- plex. Wu et al. (1990) suggested that activators facilitate the break-
able novel binders, such as alkali-activated ground-granulated blast age of the Si-O and Al-O bonds in the GGBS and enable the
formation of hydration products. Song et al. (2000) found that the
1
Ph.D. Student, Institute of Geotechnical Engineering, Southeast Univ., hydration rate of NaOH-activated GGBS depended on the pH of
Nanjing 210096, China; presently, Postdoctoral Fellow, Dept. of Civil and the starting solution. They further suggested that the pH of the mix-
Environmental Engineering, Univ. of Alberta, Edmonton, AB, Canada T6G ing solution should be higher than 11.5 to effectively activate the
2W2 (corresponding author). E-mail: yaolin@ualberta.ca hydration of GGBS. Song et al. (2000) summarized that the main
2
Graduate Student, Institute of Geotechnical Engineering, Southeast hydration product of GGBS was CSH with a low Ca=Si ratio, re-
Univ., Nanjing 210096, China. E-mail: lichen0907nj@163.com gardless of the activator used, but the morphology and composition
3
Professor, Institute of Geotechnical Engineering, Southeast Univ., of the hydration products change with the activator type and other
Nanjing 210096, China. E-mail: liusy@seu.edu.cn
hydration conditions.
Note. This manuscript was submitted on January 13, 2014; approved on
April 7, 2014; published online on July 24, 2014. Discussion period open For soil stabilization, the most commonly used GGBS activator
until December 24, 2014; separate discussions must be submitted for in- is, however, lime [CaO or CaðOHÞ2 ] (Higgins 2005; Nidzam and
dividual papers. This paper is part of the Journal of Materials in Civil Kinuthia 2010). Lime-GGBS has been used for soil-stabilization
Engineering, © ASCE, ISSN 0899-1561/04014146(7)/$25.00. applications for more than 10 years, and most of the applications

© ASCE 04014146-1 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2015, 27(4): 04014146


deal with suppressing the expansion associated with the presence of
sulfates or sulfides in lime-stabilized soil by partially substituting
GGBS with lime (Wild et al. 1996, 1998, 1999; Tasong et al. 1999;
Higgins 2005). These studies illustrated that a proper substitution
of GGBS with lime could reduce the expansion in sulphate-bearing
soils. Recently, the use of lime-GGBS for clay stabilization in the
production of unfired bricks (Oti et al. 2008, 2009a, b, c, 2010;
Kinuthia and Oti 2012) and for overcoming the deleterious effect
of flooding (Obuzor et al. 2011a, b, 2012) has also produced prom-
ising results. Soft ground improvement is another application for
Downloaded from ascelibrary.org by Sardar Vallabhbhai Natl Inst Of Tech (svnit) on 07/27/16. Copyright ASCE. For personal use only; all rights reserved.

lime-GGBS soil stabilization; Yi et al. (2014a) investigated the


lime-GGBS used for a marine soft-clay stabilization compared with
PC, and found that the optimum 90-day unconfined compressive
strength (UCS) of lime-GGBS-stabilized clay was 1.7 times that
of the PC-stabilized clay. Nidzam and Kinuthia (2010) stated that
the hydration products of lime-GGBS were similar to those of PC,
such as CSH, CAH, CASH, and ettringite. James et al. (2008) de-
tected CSH, CASH, and hydrotalcite in lime-GGBS-stabilized clay
Fig. 1. Particle-size distribution curve of the clay from laser diffrac-
through X-ray diffraction (XRD). Oti et al. (2009a) found CSH gel
tometry analysis
formed in the lime-GGBS-stabilized clay based on scanning elec-
tron microscopy (SEM) with a solid-state backscattered detection
and energy dispersive X-ray analyses. Yi et al. (2014b) revealed
that the main hydration product of lime-GGBS paste was CSH,
The water content used to prepare the stabilized clay was 90%,
along with alumino-ferrite monosulfate (AFm) through extensive
which was higher than the in situ water content considering the
microstructural investigations including XRD, SEM, Fourier trans-
additional water needed to produce binder slurry using the deep
form infrared spectroscopy, energy dispersive X-ray, and thermo
mixing method.
gravimetric analysis.
GGBS, NaOH, Na2 CO3 , CS, Na2 SO4 , and PC [PC 32.5 accord-
Lime used in civil engineering applications is generally calcined
ing to CBMA (2008)] were obtained from Nanjing, Jiangsu, China.
from limestone (CaCO3 ), which also produces considerable CO2
NaOH, Na2 CO3 , and Na2 SO4 were chemically pure. The main
emissions, requires significant energy consumption, and uses non-
chemical properties of GGBS, CS, and PC are provided in Table 1.
renewable resources. Carbide slag (CS) is an industrial waste of
The 67.98% of CaO content in the CS is equal to 89.83% of
acetylene gas production, and is composed of CaðOHÞ2 (85–95%)
with minor contents of CaCO3 (1–10%), carbon, and silicates CaðOHÞ2 ; however, the CaðOHÞ2 content in the CS is lower than
(1–3%) (Cardoso et al. 2009). In China, dry CS is generally land- that value because of the presence of CaCO3 (Cardoso et al. 2009).
filled outside the chlor-alkali plants, resulting in land occupation, Seven binders were used in this study, which were NaOH-activated
waste of calcium resources, and environmental pollution as a result GGBS (NaOH-GGBS), Na2 CO3 -activated GGBS (Na2 CO3 -
of the high alkalinity of CS, although not classified as dangerous/ GGBS), CS-activated GGBS (CS-GGBS), NaOH-CS-activated
hazardous (Cardoso et al. 2009). Because of their very similar GGBS (NaOH-CS-GGBS), Na2 CO3 -CS-activated GGBS (Na2 CO3 -
chemical properties, CS has the potential to replace lime as a build- CS-GGBS), Na2 SO4 -CS-activated GGBS (Na2 SO4 -CS-GGBS),
ing or construction material (Cardoso et al. 2009). and PC. The binder content, in terms of the weight of GGBS
To reduce the use of PC in soft-clay stabilization, this study in- or PC over the weight of dry clay, of 30% was used. According
vestigated the stabilization efficacy of alkali-activated GGBS for a to Yi et al. (2014a), hydrated lime/GGBS of 10% was the opti-
marine soft clay compared with that of PC (control). The influence mum ratio for yielding the highest 90-day UCS of stabilized
of activators, including NaOH, Na2 CO3 , CS, NaOH-CS, Na2 CO3 - clay; hence, in this study, CS/GGBS of 10% was used for CS-
CS, and Na2 SO4 -CS, on the stabilization efficacy was investigated. GGBS. According to Shi et al. (2006), NaOH solubility of 5%
A range of tests was conducted to investigate the properties of (1.25 mol=L) was used for NaOH-GGBS, and the same Naþ
the stabilized clays, including UCS, XRD, SEM, and mercury molar solubility was used for Na2 CO3 -GGBS. For activator-
intrusion porosimetry (MIP). CS-GGBS, only half Naþ molar solubility (0.625 mol=L) was
used, because CS was an alkali and its addition was equal to
approximately 0.5 mol=L NaOH in terms of OH− solubility. The
Materials and Methods activator dosages for the six types of alkali-activated GGBS are
presented in Table 2. The water in the clay (90%) was used to
the dissolve the activator before mixing with dry clay.
Marine Soft Clay and Binders
The marine soft clay was obtained approximately 2 m below
ground surface at the construction site of Dongshugang highway,
Lianyungang, Jiangsu Province, China. The clay had a plastic limit Table 1. Main Chemical Composition (% by Weight) of GGBS, CS,
of 33%, liquid limit of 74%, and water content of 75–81%. The and PC
specific gravity of the clay was 2.62, the bulk density was approx-
Loss on
imately 1.6 g=cm3, and the void ratio was approximately 1.94. The
Material CaO SiO2 Al2 O3 SO3 Fe2 O3 MgO K2 O TiO2 ignition
undrained shear strength, determined by in situ vane shear test, was
approximately 15 kPa. The particle-size-distribution curve of the GGBS 34.00 34.30 17.90 1.64 1.02 6.02 0.64 1.17 2.67
clay from laser diffractometry analysis (Mastersizer Micro MAF CS 67.98 4.01 2.30 0.32 0.13 0.27 <0.01 0.05 24.80
PC 48.80 27.40 11.50 3.28 3.43 1.16 1.31 0.48 2.00
5000, Malvern Instruments, Malvern, U.K.) is presented in Fig. 1.

© ASCE 04014146-2 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2015, 27(4): 04014146


Table 2. Activator Dosages for the Six Types of Alkali-Activated GGBS to acquire highly magnified micro images. The MIP testing was con-
Alkali-activated GGBS Activator dosage ducted by an AutoPore IV 9500 (Micromeritics Instrument Corp.,
Norcross, Georgia).
NaOH-GGBS NaOH: 1.25 mol=L
Na2 CO3 -GGBS Na2 CO3 : 0.625 mol=L
CS-GGBS CS=GGBS ¼ 0.1
NaOH-CS-GGBS NaOH: 0.625 mol=L, CS=GGBS ¼ 0.1 Results and Discussion
Na2 CO3 -CS-GGBS Na2 CO3 : 0.313 mol=L, CS=GGBS ¼ 0.1
Na2 SO4 -CS-GGBS Na2 SO4 : 0.313 mol=L, CS=GGBS ¼ 0.1 Unconfined Compressive Strength
The Na2 CO3 -GGBS stabilized-clay specimens were not strong
enough to be demolded even after 180 days of curing, although
Downloaded from ascelibrary.org by Sardar Vallabhbhai Natl Inst Of Tech (svnit) on 07/27/16. Copyright ASCE. For personal use only; all rights reserved.

Stabilized Clay Specimen Preparation


Na2 CO3 was reported as an effective activator for GGBS paste
The construction process of deep mixing method, which mixes the (Shi et al. 2006). The main reason might be that the water/binder
binder with the in situ clay using mixing blades, includes the pri- ratio in the stabilized clay was substantially higher (3.0) than that in
mary mixing effect without considerable compaction effect. Hence, paste (generally <0.5), and the Na2 CO3 hardly created a high pH
the laboratory-stabilized clay preparation procedure attempted to environment in the clay-water-GGBS system because of the low
simulate the in situ deep mixing method by using a laboratory alkalinity of Na2 CO3 . As reported by Song et al. (2000), the rate
mixer. of activating reactions depended on the pH of the starting solution.
The clay was first dried in the oven at a temperature of 105°C, For the same reason, the 7-day CS-GGBS specimens were not
and was then ground into powder and passed through 2-mm sieves. strong enough to be demolded. Yi et al. (2014a) reported that
The required amounts of raw materials (i.e., dry clay, PC, GGBS, 20% hydrated lime–GGBS stabilized clay with hydrated lime=
activators, and water) were calculated and weighed. The water, GGBS ¼ 10% had a 7-day UCS of 0.33 MPa; the same GGBS
calculated according to the clay water content (90%), was used but a different marine clay with a liquid limit of 58% and water
to the dissolve the activators before mixing with dry clay. The content of 60% were used. Because CS has a similar chemical
dry clay and the GGBS (or PC) were initially mixed and homog- composition to hydrated lime, the water content appears to have
enized for 10 min in a laboratory mixer, after which the water- a significant effect on the early-age strength development of
activator solution (or water) was added and the mixing continued alkali-activated GGBS stabilized clay. This result agrees with
for an additional 10 min. The homogenized clay-binder mixture the findings from the alkali-activated cement paste that a high
was then placed in cylindrical molds, 50 mm in diameter and water/binder ratio induces high initial porosity of sample and hence
100 mm high, applying consistent moderate compaction in three results in low strength (Shi et al. 2006).
layers for 10 min. The compaction was conducted manually by The UCS of different stabilized clays at different curing ages is
an 8-mm-diameter steel rod to eliminate air pockets, so as to im- presented in Fig. 2, with the margins of error shown by the error
prove the homogeneity of the specimens. Because the clay used in bars. NaOH-GGBS stabilized clay was found to yield the highest 7,
this study had a high water content of 90%, higher than its liquid 28, and 90-day UCS, indicating the highest activating efficacy for
limit (74%), the clay-binder mixture was close to slurry form and GGBS because of its highest alkalinity. However, significant
the rod compaction would not affect much on the sample density UCS decrease (48%) occurred from 90 to 180 days for NaOH-
except for the air-pocket-eliminating effect. The prepared specimens GGBS-stabilized clay. CS-GGBS-stabilized clay had higher 90
were placed in a sealed plastic container, where relative humidity and 180-day UCS than that of PC-stabilized clay, but significantly
and temperature were maintained at 95  3% and 20  2°C, lower 7 and 28-day UCS.
respectively. The stabilized clay samples were demolded and Because the early-age strength of stabilized clay is important for
subjected to testing after 7, 28, 90, and 180 days of curing. practice, NaOH, Na2 CO3 , and Na2 SO4 were used to increase
the strength-development rate of CS-GGBS-stabilized clay. Fig. 2
demonstrates that all three activators enhanced the 7 and 28-day
Testing Procedure UCS of CS-stabilized clay. As expected, the 7, 28, and 90-day
The diameter, height, and weight of the demolded specimens were UCS of NaOH-CS-GGBS-stabilized clay are higher than that of
measured, and then the bulk densities of the specimens were cal-
culated. The specimens with significantly high variation in bulk
density were eliminated from testing. The UCS was tested accord-
ing to ASTM (2007) in triplicate with the vertical load applied at a
constant displacement rate of 1 mm=min until failure from which
the strength was calculated. Microstructural analyses were con-
ducted by employing XRD, SEM, and MIP on the selected
mixtures taken from the center of the crushed UCS samples.
The crushed UCS samples were soaked in ethanol for 7 days to
stop the hydration reactions, and were then frozen by liquid nitro-
gen for freeze-drying. After that, the samples were placed in a vac-
uum to sublimate for 48 h. The dried sample pieces not exceeding
10 mm in size were used for SEM and MIP testing, and the ground
sample powder, sieved through 75-μm sieves, was used for
XRD testing. The XRD testing was performed using a powder
diffractometer D8 Discover (Bruker Corporation, Billerica,
Massachusetts) with a Cu Kα source to identify the crystalline
phases within the cement-stabilized matrix. A SM-6300 scanning
Fig. 2. UCS of stabilized clays at 7, 28, 90, and 180 days
electron microscope (JEOL, Akishima-Shi, Japan) was employed

© ASCE 04014146-3 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2015, 27(4): 04014146


the CS-GGBS-stabilized clay and lower than that of the NaOH-
GGBS-stabilized clay. However, considerable strength decrease
from 90 to 180 days also occurred for NaOH-CS-GGBS-stabilized
clay. Although Na2 CO3 -GGBS had no stabilizing efficacy for this
clay, Na2 CO3 increased the 7, 28, and 90-day UCS of CS-GGBS-
stabilized clay, as shown in Fig. 2. This effect is attributed to the
production of NaOH from the reaction of CS and Na2 CO3 , as
shown in Eq. (1). For the same reason, strength decreased from
90 to 180 days for Na2 CO3 -CS-GGBS-stabilized clay. The 7 and
180-day UCS of Na2 SO4 -CS-GGBS-stabilized clay was twice that
Downloaded from ascelibrary.org by Sardar Vallabhbhai Natl Inst Of Tech (svnit) on 07/27/16. Copyright ASCE. For personal use only; all rights reserved.

of the NaOH-CS-GGBS-stabilized clay, although the 28 and


90-day UCS of the former is slightly lower than the latter. More
importantly, the UCS of Na2 SO4 -CS-GGBS-stabilized clay in-
creased 25% from 90 to 180 days. Na2 SO4 -CS-GGBS-stabilized
clay yielded at least twice higher UCS than that of the PC-stabilized
clay at any curing age investigated. This result also indicates good
resistance of CS-GGBS to Na2 SO4 attack, which is important for
the durability of stabilized marine clay. There are two reasons for
the effect of Na2 SO4 on the CS-GGBS. First, NaOH resulted from
the reaction of CS and Na2 SO4 as shown in Eq. (2), which in-
creased the 7, 28, and 90-day UCS of CS-GGBS-stabilized clay
but could also decrease the 180-day UCS. Second, CaSO4 was
produced according to Eq. (2), which participated in the hydration
reactions of GGBS and changed the hydration products.

Na2 CO3 þ CaðOHÞ2 ¼ CaCO3 ↓ þ 2NaOH ð1Þ

Na2 SO4 þ CaðOHÞ2 ¼ CaSO4 ↓ þ 2NaOH ð2Þ

Fig. 3. XRD diffractograms of NaOH-GGBS, CS-GGBS, and


X-Ray Diffraction Na2 SO4 -CS-GGBS-stabilized clays at 90 and 180 days
The crystalline phases, determined by XRD analysis, are shown in
Fig. 3 for NaOH-GGBS, CS-GGBS, and Na2 SO4 -CS-GGBS-
stabilized clays at 90 and 180 days. Quartz (Q), calcite (C), phases between the 90 and 180-day stabilized clays with the same
kaolinite (K), illite (I), and clinochlore (Cl) were detected in all binder.
of the stabilized clays, reflecting the nature of the used marine clay.
CSH was also detected in all of the stabilized clays, indicating that
Scanning Electron Microscopy
the main hydration product of the three alkali-activated GGBS was
similar to that of PC. This observation is consistent with findings Typical SEM micrographs for NaOH-GGBS, CS-GGBS, and
from previous studies that the main hydration product of GGBS Na2 SO4 -CS-GGBS-stabilized clays at 90 and 180 days are shown
was CSH, regardless of the activator type (Song et al. 2000; Shi in Fig. 4. It is evident that amorphous and gel-like CSH appears in
et al. 2006; James et al. 2008; Oti et al. 2009a; Nidzam and all of the activated GGBS-stabilized clays [Figs. 4(a–f)], confirm-
Kinuthia 2010; Yi et al. 2014a, b, c). In addition to CSH, calcium ing that CSH is their common hydration product. For NaOH-
aluminate hydrates (CAH) and hydrocalumite (H) were produced GGBS-stabilized clays, reticular/honey-combed microstructures
in the NaOH-GGBS and CS-GGBS-stabilized clays. Hydro- formed by fibrillar CSH are also shown in their SEM images
calumite belongs to a family of hydrated calcium aluminates [Figs. 4(a and b)]. Compared with NaOH-GGBS-stabilized clays,
and generally refers to AFm in PC-hydration products (Matschei CS-GGBS-stabilized clays exhibited denser microstructures as
et al. 2007). The production of CAH and AFm during NaOH- shown in Figs. 4(c and d), and reticular/honey-combed CSH is
GGBS and CaðOHÞ2 -GGBS hydration was also reported in absent. Their different morphology might be attributed to their dif-
previous studies (Shi et al. 2006; Yi et al. 2014b). Although the ferent Ca=Si ratio of CSH, as presented in the XRD section. Addi-
hydration products of NaOH-GGBS and CS-GGBS-stabilized tionally, small-sized, needle-like ettringite appears in Figs. 4(c and
clays were similar, the Ca=Si ratio of CSH might be higher in d), consistent with the XRD results. Evidently, plenty of needle-
the latter than in the former because of the addition of Ca2þ from like ettringite of different sizes appears in the SEM images of
CS. Additionally, a weak peak of ettringite (E) displays in the XRD Na2 SO4 -CS-GGBS-stabilized clays [Figs. 4(e and f)]. CSH, with
diffractograms of the CS-GGBS-stabilized clay at 90 or 180 days. a similar morphology as that of the CS-GGBS-stabilized clays
The potential formation of ettringite during CaðOHÞ2 -GGBS hy- [Figs. 4(c and d)], also appears in the Na2 SO4 -CS-GGBS-stabilized
dration was previously stated by Nidzam and Kinuthia (2010). clays. From the SEM images in Fig. 4, no significant difference is
After the addition of Na2 SO4 into the CS-GGBS-stabilized clays, found in the hydration products or morphology between the 90 and
CAH and hydrocalumite disappeared; instead, more ettringite pro- 180-day stabilized clay with the same binder.
duced. The change in hydration products was attributed to the
CaSO4 that produced according to Eq. (2), because CaSO4 could
Mercury Intrusion Porosimetry
enhance the ettringite production and convert CAH and AFm to
ettringite (Irassar et al. 2003; Tixier and Mobasher, 2003). Fig. 3 Figs. 5 and 6 present the mercury intrusion curves and differential
also shows that there is no considerable difference in crystalline pore-size distributions, respectively, for NaOH-GGBS, CS-GGBS,

© ASCE 04014146-4 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2015, 27(4): 04014146


Downloaded from ascelibrary.org by Sardar Vallabhbhai Natl Inst Of Tech (svnit) on 07/27/16. Copyright ASCE. For personal use only; all rights reserved.

Fig. 4. Typical scanning electron micrographs of stabilized clays: (a) 90-day NaOH-GGBS; (b) 180-day NaOH-GGBS; (c) 90-day CS-GGBS;
(d) 180-day CS-GGBS; (e) 90-day Na2 SO4 -CS-GGBS; (f) 180-day Na2 SO4 -CS-GGBS

and Na2 SO4 -CS-GGBS-stabilized clays at 90 and 180 days. MIP free water in the clay and filled its porosity, which enhanced the
was conducted with an entrance pore radius from 0.007 to 340 μm; strength of stabilized clay. However, excessive ettringite would also
however, the results with an entrance pore radius greater than cause cracks in the stabilized soil, thus reducing its strength
10 μm are not presented in Figs. 5 and 6, as they are indentical (Rajasekaran 2005), especially when the initial stabilized soil
for all of the samples. Fig. 5 shows that the CS-GGBS-stabilized has a low porosity. Fig. 5 shows that the porosities of both CS-
clay has the highest 90-day total intrusion volume (total porosity), GGBS and Na2 SO4 -CS-GGBS-stabilized clays decreased from
because of the low hydration rate of CS-GGBS as indicated by the 90 to 180 days, because of the hydration products produced from
UCS results. The Na2 SO4 -CS-GGBS-stabilized clay yielded al- the residual GGBS, which also resulted in the UCS increase shown
most the same 90-day porosity as that of NaOH-GGBS-stabilized in Fig. 2. On the contrary, the porosity of NaOH-GGBS-stabilized
clay, although the 90-day UCS of the former was lower than the clay increased from 90 to 180 days, corresponding to its UCS
latter. This result might be attributed to the following reasons. decrease (Fig. 2).
NaOH-GGBS had the highest hydration rate as a result of the high- Fig. 6 shows the differential pore-size distributions derived from
est alkalinity of NaOH, and produced the most CSH, the major Fig. 5. The CS-GGBS-stabilized clay had more large-sized pores
binding product among the hydration products. For the 90-day (0.5–2 μm) than the other two stabilized clays at the same age.
Na2 SO4 -CS-GGBS-stabilized clay, less CSH was produced Compared with the CS-GGBS-stabilized clay at the same age,
than that in the 90-day NaOH-GGBS-stabilized clay, resulting the Na2 SO4 -CS-GGBS-stabilized clay had negligible pores with
in lower strength. In contrast, the formation of ettringite a diameter greater than 0.6 μm, but more small-sized pores with
(3CaO · Al2 O3 · 3CaSO4 · 32H2 O) stabilized a large quantity of a diameter ranging from 0.1 to 0.4 μm. As discussed previously,

© ASCE 04014146-5 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2015, 27(4): 04014146


pore change are not clear from the tests conducted in this study, be-
cause no significant change was observed in XRD and SEM results
as presented previously. Collins and Sanjayan (2001) reported that
the lack of moist curing of alkali-activated slag concrete increased
the level of micro cracking and resulted in a strength decrease after
91 days. In Collins and Sanjayan (2001), the activator was a mixture
of sodium silicate and hydrated lime, the samples cured at 50% rel-
ative humidity and 23°C, and the microcracking was investigated by
using both MIP and a crack-detection microscope. Nevertheless, fur-
ther study is needed to better understand the strength decrease for the
Downloaded from ascelibrary.org by Sardar Vallabhbhai Natl Inst Of Tech (svnit) on 07/27/16. Copyright ASCE. For personal use only; all rights reserved.

NaOH-GGBS-stabilized clay from 90 to 180 days.

Conclusions

Na2 CO3 -GGBS had no stabilization efficacy for the marine soft
clay used in this study. NaOH-GGBS-stabilized clay had the high-
est UCS at ages 7, 28, and 90 days; however, the strength of the
Fig. 5. Mercury intrusion curves of NaOH-GGBS, CS-GGBS, and stabilized clay decreased from 90 to 180 days. CS-GGBS-stabilized
Na2 SO4 -CS-GGBS-stabilized clays at 90 and 180 days clay had higher 90 and 180-day UCS than that of PC-stabilized
clay, but significantly lower 7 and 28-day UCS.
The addition of NaOH, Na2 CO3 , and Na2 SO4 increased the
this pore change was primarily attributed to the formation of ettrin- strength development rate of CS-GGBS-stabilized clay. However,
gite, which stabilized free water, filled the large-size pores in the the UCS of NaOH-CS-GGBS-stabilized clay decreased from 90
clay, and converted them to relatively small-sized pores. The to 180 days as well. The effect of Na2 CO3 on CS-GGBS was
NaOH-GGBS-stabilized clays had fewer large-sized pores than through the production of NaOH from the reaction between
the other two types of stabilized clays, but more small-sized pores Na2 CO3 and CS, and a similar strength decrease occurred
with a diameter ranging from 0.06 to 0.1 μm. This resulted because for Na2 CO3 -CS-GGBS-stabilized clay from 90 to 180 days.
NaOH had the highest activating effect and produced the most Na2 SO4 -CS-GGBS was found to be the optimum binder for stabi-
small-sized CSH, which filled most of the large-sized pores in lizing this soft clay, yielding at least twice higher UCS than that of
the stabilized clay, and converted them to small-size pores. From PC-stabilized clay at any age studied. This result indicated good
90 to 180 days there was a reduction in large-sized pores and an resistance of CS-GGBS-stabilized clay to Na2 SO4 attack.
increase in relatively small-sized pores in both CS-GGBS and XRD and SEM results indicated that the hydration products for
Na2 SO4 -CS-GGBS-stabilized clays, as shown in Fig. 6. However, NaOH-GGBS-stabilized clays included CSH, CAH, and AFm,
the NaOH-GGBS-stabilized clays exhibited an opposite trend, which were similar to those of the CS-GGBS-stabilized clay. MIP
i.e., an increase of pores with relatively large-sized diameter rang- results revealed that microcracking and increasing porosity occurred
ing from 0.08 to 0.5 μm and a decrease of relatively small-sized in the NaOH-GGBS-stabilized clay from 90 to 180 days, which
pores with a diameter ranging from 0.02 to 0.07 μm. Furthermore, might be the main reason for its strength decrease. More investigation
Fig. 5 shows that the total pore volume of Na2 SO4 -CS-GGBS- is required to further understand the strength-loss mechanism.
stabilized clay increased from 90 to 180 days. These unusual pore In addition to the NaOH resulting from the reaction between
changes of NaOH-GGBS-stabilized clay indicated that microcrack- Na2 SO4 and CS, CaSO4 was also produced, which contributed
ing occurred in stabilized clay from 90 to 180 days, which deceased to the production of ettringtite and the absence of CAH and
the strength of stabilized clay. However, the reasons that caused this AFm. The formation of ettringtite could stabilize free water in the
clay, reduce its porosity, and enhance its strength. However, exces-
sive ettringite would also reduce its strength. Because the main
mechanism of Na2 SO4 to improve the efficacy of CS-GGBS for sta-
bilized clay is through CaSO4 , the use of Na2 SO4 might be replaced
with phosphogypsum, an industry waste whose main composition is
CaSO4 , for environmental and economical consideration.

Acknowledgments

The authors greatly appreciate the funding provided by National


Science and Technology Pillar Program (2012BAJ01B02-01)
and Natural Science Foundation of Jiangsu Province
(SBK2014040468). The authors thank Dr. Somayeh Nassiri and
Dr. Leon F. Gay at the University of Alberta, Canada, for their
valuable comments.

References
Fig. 6. Differential pore-size distribution of NaOH-GGBS, CS-GGBS,
ASTM. (2007). “Standard method for compressive strength of molded
and Na2 SO4 -CS-GGBS-stabilized clays at 90 and 180 days
soil-cement cylinders.” D1633-00, West Conshohocken, PA.

© ASCE 04014146-6 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2015, 27(4): 04014146


Bakharev, T., Sanjayan, J. G., and Cheng, Y. B. (1999). “Alkali activation of Oti, J. E., Kinuthia, J. M., and Bai, J. (2009c). “Unfired clay bricks:
Australian slag cements.” Cem. Concr. Res., 29(1), 113–120. From laboratory to industrial production.” Eng. Sustain., 162(4),
Bergado, D. T., Anderson, L. R., Miura, N., and Balasubramaniam, A. S. 229–237.
(1996). Soft ground improvement in lowland and other environments, Oti, J. E., Kinuthia, J. M., and Bai, J. (2010). “Freeze-thaw of stabilised
ASCE, Reston, VA. clay brick.” Waste Resour. Manage., 163(3), 129–135.
Cardoso, F. A., Fernandes, H. C., Pileggi, R. G., Cincotto, M. A., and John, Porbaha, A. (1998). “State-of-the-art in deep mixing technology. Part I:
V. M. (2009). “Carbide lime and industrial hydrated lime characteriza- Basic concepts and overview of technology.” Ground Improv., 2(2),
tion.” Powder Technol., 195(2), 143–149. 81–92.
Chew, S. H., Kamruzzaman, A. H. M., and Lee, F. H. (2004). “Physico- Porbaha, A., Shibuya, S., and Kishida, T. (2000). “State of the art in deep
chemical and engineering behavior of cement treated clays.” J. Geotech. mixing technology. Part III: Geomaterial characterization.” Ground
Geoenviron. Eng., 10.1061/(ASCE)1090-0241(2004)130:7(696), Improv., 4(3), 91–100.
696–706. Rajasekaran, G. (2005). “Sulphate attack and ettringite formation in the
Downloaded from ascelibrary.org by Sardar Vallabhbhai Natl Inst Of Tech (svnit) on 07/27/16. Copyright ASCE. For personal use only; all rights reserved.

China Building Materials Academy (CBMA). (2008). Common portland lime and cement stabilized marine clays.” Ocean Eng., 32(8–9),
cement (GB175-2007), China Standard Press, Beijing (in Chinese). 1133–1159.
Collins, F., and Sanjayan, J. G. (2001). “Microcracking and strength Roy, D. M. (1999). “Alkali-activated cements: Opportunities and
development of alkali activated slag concrete.” Cem. Concr. Compos., challenges.” Cem. Concr. Res., 29(2), 249–254.
23(4–5), 345–352. Scrivener, K. L., and Kirkpatrick, R. J. (2008). “Innovation in use and re-
Higgins, D. D. (2005). Soil stabilisation with ground granulated blastfur- search on cementitious material.” Cem. Concr. Res., 38(2), 128–136.
nace slag, U.K. Cementitious Slag makers Association, Surrey, BC, Shi, C., Krivenko, P. V., and Roy, D. (2006). Alkali-activated cements and
Canada. concretes, Taylor and Francis, London.
Higgins, D. D. (2007). “GGBS and sustainability.” Constr. Mater., 160(3), Song, S., Sohn, D., Jennings, H. M., and Mason, T. O. (2000). “Hydration
99–101. of alkali-activated ground granulated blastfurnace slag.” J. Mater. Sci.,
Holm, G. (2003). “State of practice in dry deep mixing methods.” Proc., 35(1), 249–257.
3rd Int. Specialty Conf. on Grouting and Ground Treatment, ASCE, Tasong, W. A., Wild, S., and Tilley, R. J. D. (1999). “Mechanism by which
Reston, VA, 145–163. ground granulated blastfurnace slag prevents sulphate attack of
Irassar, E. F., Bonavetti, V. L., and Gonzalez, M. (2003). “Microstructural lime-stabilised kaolinite.” Cem. Concr. Res., 29(7), 975–982.
study of sulfate attack on ordinary and limestone portland cements at Terashi, M. (2003). “The state of practice in deep mixing methods.” Proc.,
ambient temperature.” Cem. Concr. Res., 33(1), 31–41. 3rd Int. Specialty Conf. on Grouting and Ground Treatment, ASCE,
James, R., Kamruzzaman, A. H. M., Haqueand, A., and Wilkinson, A. Reston, VA, 25–49.
(2008). “Behaviour of lime-slag-treated clay.” Ground Improv., Tixier, R., and Mobasher, B. (2003). “Modeling of damage in cement-based
161(4), 207–221. materials subjected to external sulphate attack. I: Formulation.”
Kinuthia, J. M., and Oti, J. E. (2012). “Designed non-fired clay mixes for J. Mater. Civ. Eng., 10.1061/(ASCE)0899-1561(2003)15:4(305),
sustainable and low carbon use.” Appl. Clay Sci., 59–60, 131–139. 305–313.
Kitazume, M., and Terash, M. (2013). The deep mixing method, CRC Press/ Wang, S. D., and Scrivener, K. L. (1995). “Hydration products of alkali
Balkema, Leiden, Netherlands. activated slag cement.” Cem. Concr. Res., 25(3), 561–571.
Matschei, T., Lothenbach, B., and Gasser, F. P. (2007). “The AFm phase in Wild, S., Kinuthia, J. M., Jones, G. I., and Higgins, D. D. (1998). “Effects
portland cement.” Cem. Concr. Res., 37(2), 118–130. of partial substitution of lime with ground granulated blast furnace slag
Nidzam, R. M., and Kinuthia, J. M. (2010). “Sustainable soil stabilisation (GGBS) on the strength properties of lime stabilised sulphate-bearing
with blastfurnace slag—A review.” Constr. Mater., 163(3), 157–165. clay soils.” Eng. Geol., 51(1), 37–53.
Obuzor, G. N., Kinuthia, J. M., and Robinson, R. B. (2011a). “Enhancing Wild, S., Kinuthia, J. M., Jones, G. I., and Higgins, D. D. (1999).
the durability of flooded low-capacity soils by utilizing lime-activated “Suppression of swelling associated with ettringite formation in lime
ground granulated blastfurnace slag (GGBS).” Eng. Geol., 123(3), stabilized sulphate bearing clay soils by partial substitution of lime with
179–186. granulated blastfurnace slag.” Eng. Geol., 51(4), 257–277.
Obuzor, G. N., Kinuthia, J. M., and Robinson, R. B. (2011b). “Utilisation Wild, S., Kinuthia, J. M., Robinson, R. B., and Humphreys, I. (1996).
of lime activated GGBS to reduce the deleterious effect of flooding on “Effects of ground granulated blastfurnace slag (GGBS) on strength
stabilised road structural materials: A laboratory simulation.” Eng. and swelling properties of lime stabilised kaolinite in the presence
Geol., 122(3-4), 334–338. of sulphates.” Clay Miner., 31(3), 423–433.
Obuzor, G. N., Kinuthia, J. M., and Robinson, R. B. (2012). “Soil stabi- Wu, X., Jiang, W., and Roy, D. M. (1990). “Early activation and properties
lisation with lime-activated-GGBS—A mitigation to flooding effects on of slag cement.” Cem. Concr. Res., 20(6), 961–974.
road structural layers/embankments constructed on floodplains.” Eng. Yi, Y., Gu, L., and Liu, S. (2014a). “Mechanical and microstructural prop-
Geol., 151, 112–119. erties of lime-activated GGBS stabilized marine soft clay.” Appl. Clay
Oti, J. E., Kinuthia, J. M., and Bai, J. (2008). “Using slag for unfired clay Sci., in press.
masonry-bricks.” Constr. Mater., 161(4), 147–155. Yi, Y., Liska, M., and Al-Tabbaa, A. (2014b). “Properties and microstruc-
Oti, J. E., Kinuthia, J. M., and Bai, J. (2009a). “Compressive strength and ture of GGBS-MgO pastes.” Adv. Cem. Res., 26(2), 114–122.
microstructural analysis of unfired clay masonry bricks.” Eng. Geol., Yi, Y., Liska, M., and Al-Tabbaa, A. (2014c). “Properties of two model
109(3–4), 230–240. soils stabilised with different blends and contents of GGBS, MgO,
Oti, J. E., Kinuthia, J. M., and Bai, J. (2009b). “Engineering properties of lime and PC.” J. Mater. Civ. Eng., 10.1061/(ASCE)MT.1943-5533
unfired clay masonry bricks.” Eng. Geol., 107(3–4), 130–139. .0000806, 267–274.

© ASCE 04014146-7 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2015, 27(4): 04014146

Das könnte Ihnen auch gefallen