Johnston and Santagata 2011 - 1D Consolidation Behaviour of Cement-Treated Soil

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

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

net/publication/237373288

One-dimensional consolidation behavior of cement-treated organic soil

Article  in  Canadian Geotechnical Journal · July 2011


DOI: 10.1139/t11-020

CITATIONS READS
46 555

4 authors, including:

Cliff T. Johnston Marika Santagata


Purdue University Haute école de santé Genève
230 PUBLICATIONS   9,477 CITATIONS    28 PUBLICATIONS   542 CITATIONS   

SEE PROFILE SEE PROFILE

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

Mineral-Water Interaction View project

The nature and properties of mineral and organic matter associations in soils View project

All content following this page was uploaded by Cliff T. Johnston on 12 August 2015.

The user has requested enhancement of the downloaded file.


1100

One-dimensional consolidation behavior of


cement-treated organic soil
Antonio Bobet, Joonho Hwang, Cliff T. Johnston, and Marika Santagata

Abstract: This paper addresses the effects of cement treatment on the one-dimensional (1D) consolidation behavior of a
highly organic soil (LOI ∼ 40%–60%, where LOI is the loss on ignition), based on 1D constant rate of strain and incremen-
tal loading tests. The effects of Portland cement addition are evaluated for dosages ranging from 8% to 100% by dry mass
of soil, corresponding to values of the cement factor of 24 and 296 kg of cement per cubic metre of untreated soil, within
the range used in deep mixing practice. Additional parameters investigated are the impact of curing surcharge and duration.
The most evident effect of the treatment is the development of a cementation-induced preconsolidation stress: the greater the
cement dosage, the greater the preconsolidation stress and the greater the vertical effective stress that can be sustained at
any void ratio. The results also provide a consistent picture of the effects of cement treatment on stiffness, hydraulic conduc-
tivity, coefficient of consolidation, and creep. Comparison to data obtained for the untreated soil demonstrates the “stable”
nature of the structure generated as a result of treatment. The consolidation results are complemented by pH measurements,
extraction tests, elemental analyses, and Fourier transform infra-red (FTIR) spectroscopy analyses, which provide insight
into the interaction between soil organic matter and cement.
Key words: peat, organic matter, cement treatment, deep mixing, consolidation, creep.
Résumé : Cet article discute des effets du traitement au ciment d’un sol hautement organique (LOI ∼ 40–60 %, où « LOI »
est la perte au feu) sur son comportement en consolidation unidimensionnelle (1D), à partir d’essais en chargement en 1D à
taux de déformation constant et incrémental. Les effets de l’ajout de ciment Portland sont évalués pour des dosages variant
de 8 à 100 %, ce qui correspond à des valeurs de facteur de ciment de 24 et 296 kg de ciment par mètre cube de sol non
traité, ceci étant des valeurs typiquement utilisées dans le mélange en profondeur. Les paramètres additionnels qui sont étu-
diés sont l’impact de la surcharge durant le curage et sa durée. L’effet le plus évident du traitement est le développement
d’une contrainte de pré-consolidation induite par la cimentation : plus le dosage de ciment est élevée, plus la contrainte de
pré-consolidation est élevée, et une plus grande contrainte effective verticale peut être soutenue pour tout indice des vides.
Les résultats fournissent aussi une image des effets du traitement au ciment sur la rigidité, la conductivité hydraulique, le co-
efficient de consolidation, et le fluage. Les résultats obtenus sont comparés à des données provenant du sol non traité, ce
qui démontre la nature « stable » de la structure générée suite au traitement. Les résultats de consolidation sont complémen-
tés par des mesures de pH, des essais d’extraction, des analyses élémentaires et des analyses de spectroscopie infrarouge à
transformée de Fourier (« FTIR »). Ces analyses fournissent des informations sur l’interaction entre la matière organique du
sol et le ciment.
Mots‐clés : tourbe, matière organique, traitement au ciment, mélange en profondeur, consolidation, fluage.
[Traduit par la Rédaction]

Introduction tions are generally available to modify and improve the


Peats and other soils high in organic matter are prone to ground conditions: strengthening of the foundation, elimina-
large settlements, which derive both from their high compres- tion of the problem soils, treatment of the problem soils, and
sibility and the significant impact of secondary consolidation. relocation of the project. In many cases, the only option is
As a result, construction on these soils poses significant chal- strengthening the foundation or elimination of part of the
lenges to the geotechnical engineering profession. Several op- problem soils, as the other options are impractical or too ex-
pensive. A widely used approach to the improvement of the
Received 28 May 2010. Accepted 17 February 2011. Published engineering properties of soft soils is the deep mixing
at www.nrcresearchpress.com/cgj on 12 July 2011. method (DMM), which was developed in Sweden and Japan
A. Bobet and M. Santagata. School of Civil Engineering, in the late 1960s for treatment of soft soils. It was introduced
Purdue University, 550 Stadium Mall Drive, West Lafayette, IN in North America in the late 1980s and has been widely used
47907, USA. since. Although still limited compared to Europe, there have
J. Hwang. Asian Development Bank, 6 ADB Avenue, been applications of this method in North America to treat-
Mandaluyong City 1550, Philippines. ment of organic soils; for example, the stabilization of a 3.5
C.T. Johnston. Department of Agronomy, Purdue University, to 5 m thick organic silty clay deposit for construction of a
915 W State Street, West Lafayette, IN 47907, USA. railroad embankment in a section of the Hudson–Bergen
Corresponding author: Marika Santagata (e-mail: mks@purdue. Light Rail Transit System, New Jersey (Esrig et al. 2003),
edu). and the stabilization of a 2.5 to 7.5 m thick organic clay layer

Can. Geotech. J. 48: 1100–1115 (2011) doi:10.1139/T11-020 Published by NRC Research Press
Bobet et al. 1101

(with organic content from 4% to over 30%) for the I-95 high- Experimental methods
way widening in Alexandria, Va. (Lambrechts et al. 2003).
The principle of deep mixing is the mixing of hardening Materials
agents (generally lime or cement) with soil in situ in con- Soil for this research program was sampled at the south end
trolled proportions to produce columns of hardened material of the “Celery Bog” park in West Lafayette, In. Two block
that display higher strength and stiffness and generally lower samples and several disturbed soil samples were collected
hydraulic conductivity than the original soil. Depending on in April and July 2001 from the bottom of a 4.5 m diameter
the configuration of the deep mixing system used, the binder pit excavated to a depth of 2 m. Seven (samples S5–S11) of
agents can be introduced in slurry (wet method) or dry (dry the 11 disturbed samples collected were used for this study,
method) form. The binding agents are mixed with soil at the and are referenced in this paper based on the sample num-
distal end or along the drill shaft by pure rotation of the mix- ber. Borehole data (Earth Exploration, Inc. 1993) indicate
ing augers or by a combination of rotation and injection of that the layer of organic soil from which the samples were
the binder in slurry form at high pressure (Bruce and Bruce obtained is 3–4 m thick and underlain by marl and a silty
2003). The properties of deep mixed soils are controlled by a clay layer, and that the groundwater table lies in the upper
variety of factors including: (i) physical and chemical proper- part of the organic soil layer. The surface material at this
ties of the soil and groundwater, (ii) type and amount of location is designated as a Houghton muck (USDA 1998).
binder, (iii) curing period, and (iv) mixing effectiveness. To The soil is a highly organic (loss on ignition (LOI) in the
date, significant research has focused on investigating the fac- 40% to 60% range), sapric (∼2% fiber content) soil with
tors influencing the improvement of the strength of inorganic mildly acidic (5.6–5.9) pH. The inorganic fraction is entirely
clays and the geotechnical properties of these soils after treat- finer than 0.075 mm, with over 60% in the clay fraction,
ment. However, despite the considerable work performed in which contains highly active clay minerals (smectite and ver-
Sweden over the years (e.g., Åhnberg and Holm 1999; miculite). The organic fraction, comprised mainly of highly
processed humic substances, functions as a “glueing” agent
Åhnberg et al. 2001; Holm 2005), the recent EuroSoilStab
between the silt-size and clay-size particles, causing the aver-
project, and additional recent contributions exploring the ef-
age size of the soil particles to increase. In contrast with
fectiveness of different binders (e.g., Hayashi and Nishimoto
nonhumic substances, which are easily attacked by soil mi-
2005; Hernandez-Martinez and Al-Tabbaa 2005, 2009) and
croorganisms and exist in the soil for a relatively short period
the durability of the treatment (e.g., Butcher 2005), there re-
of time (Schnitzer and Khan 1972; Sparks 2003), humic sub-
main many questions on the effectiveness of this method in
stances refer to naturally occurring, biogenic, heterogeneous
treating organic soils. They arise from concerns about the in- organic compounds of high molecular weight (Sparks 2003).
terference of organics with cement hydration reactions, and Humic substances, which represent one of the most chemi-
on the dissolution of organic matter in high pH environ- cally reactive fractions of the soil due to their high surface
ments. These issues are compounded by the fact that the area and surface charge, can be divided into three fractions
term “organic soils” is generally used to refer to a broad set based on their solubility characteristics: humin, humic acid,
of geomaterials, with often contrasting geotechnical proper- and fulvic acid. Humin refers to the soil organic matter frac-
ties. Moreover, assessments of the effectiveness of the treat- tion that is insoluble in alkali and remains after extraction of
ment have, to date, typically focused on the measured the humic and fulvic acids with dilute alkali; fulvic acid re-
increase in compressive strength (most commonly from un- fers to the colored soil organic matter that is soluble in both
confined tests) and, to a lesser degree (e.g., Cortellazzo and alkali and dilute acid; humic acid is the dark-colored organic
Cola 1999; Tremblay et al. 2001; Hebib and Farrell 2003), matter that is soluble in alkali, but is insoluble and precipi-
on the one-dimensional (1D) compression behavior. Studies tates in dilute acid. The fulvic fraction, in general, has lower
on other critical properties including coefficient of consolida- molecular weight, and contains higher oxygen, but lower car-
tion and creep coefficient remain very limited. bon than humic acid. The fulvic acid fraction also contains
The objective of this paper is to fill this void by presenting significantly more acidic functional groups, especially carb-
results from an extensive experimental program of constant oxyl acid (COOH), and is more reactive than humic acid.
rate of strain (CRS) and incremental loading (IL) tests con- The distribution of the three fractions of soil organic matter
ducted on a soil with approximately 50% organic content. varies depending on the soil type and depth in the soil pro-
The data presented also include chemical data that provide file. Santagata et al. (2008) report that the soil used in this
insight into the interaction of the soil’s organic matter with research is comprised of approximately 74% humin and inor-
cement. The work complements the results presented by San- ganic components, 24% humic acid, and less than 3% fulvic
tagata et al. (2008), which focused on the 1D consolidation acid.
behavior of the same organic soil in both the undisturbed (tests All index properties of the soil show significant sample to
on specimens obtained from block samples) and reconstituted sample variability with the specific gravity (Gs) exhibiting a
state. While in the field the effectiveness of the treatment linear correlation to the LOI (Gs = –0.0104LOI(%) + 2.570),
would be measured as improvement over the undisturbed with measured values of Gs ranging between 1.9 and 2.2
field conditions, the reconstituted results are used in this pa- (Santagata et al. 2008). Similarly, both the plastic limit
per as the reference against which the changes in mechani- (PL = 114%–253%) and the liquid limit (LL = 228%–406%)
cal and chemical properties of the soil treated with cement increase with increasing organic content. The latter decreased
are assessed. The choice of this baseline removes the effects to 24%–36% of its original value as a result of oven drying.
of the soil’s initial conditions, which are necessarily site de- Based on the above, the soil can be classified according to
pendent. the Unified Soil Classification System (USCS; ASTM 2006)

Published by NRC Research Press


1102 Can. Geotech. J. Vol. 48, 2011

as an organic clay with sand (OH). Details on soil index plate distributed the air pressure uniformly to the specimen,
properties, mineralogy, and microstructure are provided by reducing disturbance as the sample was extruded.
Santagata et al. (2008). At least two samples were prepared for each cement con-
Portland cement (PC) was selected as the binder as it is the tent. Specimens with a diameter of 2.5 in. (1 in. = 25.4 mm)
most economic and most widely used cement in practice. Ad- were trimmed from these samples and used to perform a pair
ditionally, and consistent with data from the literature (e.g., of constant rate of strain (CRS) and incremental loading (IL)
Hernandez-Martinez and Al-Tabbaa 2009), unconfined com- consolidation tests.
pression tests (Hwang et al. 2004) on the soil treated with Modified curing procedures were used to investigate the
different binders, including cement and lime, showed that ce- role played by curing surcharge and curing duration (see sec-
ment provided the most effective treatment. The amounts of tion titled “Effects of curing duration and curing surcharge”).
PC used ranged from 8% to 20%, 50%, and 100% by dry
mass of soil, corresponding to values of the cement factor Consolidation tests
(mass of cement used per unit volume of untreated soil) of All CRS tests presented in this paper were conducted em-
approximately 24, 59, 148, and 296 kg/m3, respectively. ploying one of three computer-controlled CRS apparatuses
These values fall in the range employed in practice (FHWA manufactured by Geotac of Houston, Tex., available in Pur-
2000; Bruce and Bruce 2003), particularly when treating due’s Bechtel Geotechnical Engineering Laboratory, and fol-
highly organic soils (e.g., Åhnberg and Holm 1999). lowing the same procedure. First the soil specimen was
backpressure saturated for 36–48 h. The value of the back
Sample preparation pressure varied depending on the cement content: from
The sample preparation procedure consisted of four steps: 300 kPa for the untreated soil to 560–600 kPa for the higher
mixing, compaction, curing, and extrusion. The mixing pro- dosage of cement. Following backpressure saturation, the
cedure was specially designed to simulate the kneading ac- specimen was loaded one-dimensionally at a constant dis-
tion applied to the soil during deep mixing. A measurement placement rate corresponding to a nominal strain rate — i.e.,
of the water content was conducted 24 h prior to mixing to based on the initial specimen height — of 0.25%–1%/h, de-
determine the exact mix proportions (at the same time a loss pending on the test, until the desired target stress was reached
of ignition test was also conducted). In the mixing phase, soil (200–1600 kPa). This load was maintained for 2–3 days until
and water were mixed at a water content of 289%, the aver- dissipation of at least 95% of the excess pore pressure gener-
age value of the natural water content measured on the April ated during loading. Then, the specimen was unloaded to
samples. Portland cement was then added to the mixture in 10% of the maximum load at 10% of the displacement rate
the form of a slurry with a water to cement ratio of 0.5, a employed during loading. The rate reduction was required to
value typical in practice. The components were then mixed limit the negative excess pore pressure generated during un-
in a 5 L Kitchen-Aid stand mixer for 2 min. loading. The load was again maintained for 2–3 days until
After mixing, the soil was compacted into plastic cylinders 95% of the negative excess pore pressure dissipated. Finally,
(diameter = 7.62 cm, height = 15.24 cm) using a modified in some cases the specimen was reloaded using the same dis-
mechanical standard Proctor hammer with a reduced diameter placement rate employed for the consolidation phase to a fi-
(2.54 cm) to generate a kneading action (Hwang et al. 2004). nal target stress. At this stage a creep test lasting several
A thin film of concrete form oil was applied to the cylinder cycles was conducted. The nonlinear solution (Wissa et al.
wall before placing the soil, to allow for easy extraction of 1971) was used for reduction of the CRS data to derive verti-
the sample. The soil mixture was compacted in three layers, cal effective stress, coefficient of consolidation (Cv), and ver-
applying 40 blows per layer. After compaction, the cylinder tical hydraulic conductivity (kv) as a continuous function of
was cut in half along the plane parallel to the ends to reduce the strain (or void ratio) from measurements of vertical dis-
friction along the walls during the curing and extrusion proc- placement, axial load, excess pore pressure developed at the
esses. A 48 kPa surcharge was applied to the soil by placing base of the soil specimen, and cell (back) pressure.
dead weights on a concrete cap positioned on top of the cyl- All incremental loading tests were performed using a cell
inder, to simulate the effect of an overburden stress. During equipped for backpressure saturation, and under single drain-
the curing period the soil cylinders were immersed in tap age conditions with measurement of the excess pore pressure
water to ensure continued access of the soil and cement to at the base of the specimen. Measurements of the base excess
water, as would occur in the field. To this effect a small pore pressure were used to determine the end of primary con-
opening was drilled in the bottom of the cylinder, where a solidation, and the load increments were applied allowing little
thin Plexiglas plate with a large number of holes was placed. to no secondary compression. This was done to limit the im-
This was done to ensure access of water to the entire cross pact of changes in soil stiffness with time in the cement-treated
section at the bottom of the specimen during curing and to soil (Kassim and Clarke 1999). The specimen was allowed to
facilitate extraction of the soil from the cylinder with mini- creep for at least one cycle of secondary compression at the
mal disturbance to the soil. Curing lasted 14 days for the ma- maximum effective stress (∼1550–1600 kPa) reached in
jority of the tests presented in this paper, although a limited each test.
number of tests were conducted on specimens cured for Table 1 summarizes the consolidation testing program. For
28 days. At the end of curing, the cylinder was removed each of the tests performed it provides the sample source
from the water bath and hot water was run over the sides to number, LOI measured on the soil prior to mixing, cement
facilitate extrusion of the soil from the mold. The cylinder content, test type, strain rate used (CRS test only), curing du-
was then inverted on a piece of wax paper and air pressure ration and surcharge, and water content (w) and void ratio (e)
was applied to the opening on its bottom. The rigid plexiglas at the start of consolidation (i.e., end of curing). As shown in

Published by NRC Research Press


Bobet et al. 1103

Table 1. Summary of experimental program.

Field Strain rate Curing duration Curing sur- w (%) end e end of
Test No. sample LOI (%) PC (%) Test type (%/h) (days) charge (kPa) of curing curing
6 S5 45.3 0 CRS 0.5 14 48 179.2 4.03
7 S5 49.2 0 CRS 0.5 14 48 197.0 4.26
8 S5 43.6 0 CRS 0.5 14 48 219.5 4.26
9 S7 54.4 0 CRS 0.25 14 48 206.6 4.32
10 S7 55.2 0 CRS 1 14 48 214.8 4.45
11 S7 54.4 0 CRS 0.1 14 48 181.9 4.20
12 S9 36.6 0 CRS 0.1 14 48 156.3 3.80
13 S9 37.5 0 CRS 1 14 48 155.4 3.64
14 S7 52.5 0 IL — 14 48 194.4 4.38
15 S8 46.0 0 IL — 14 48 178.5 3.67
16 S9 40.9 18.7 CRS 0.5 14 48 169.0 3.65
17 S9 40.7 18.7 IL — 14 48 166.9 3.74
18 S10 41.7 8.1 CRS 0.5 14 48 174.4 3.75
19 S10 41.7 8.1 IL — 14 48 180.2 3.95
20 S10 44.7 51.4 CRS 0.5 14 48 167.6 3.57
21 S10 44.7 51.4 IL — 14 48 168.0 3.57
22 S10 44.8 103.4 CRS 0.5 14 48 125.2 2.57
23 S10 44.8 103.4 IL — 14 48 126.0 2.58
24 S11 48.0 100.8 CRS 0.5 14 48 129.0 2.69
25 S10 45.0 52.6 CRS 0.5 28 48 171.1 3.65
26 S10 44.8 52.6 IL — 28 48 168.0 3.65
27 S11 42.7 100.8 CRS 0.5 28 48 130.1 2.73
28 S11 42.3 100.8 IL — 28 48 133.0 2.76
29 S10 44.2 52.6 CRS 0.5 14 96 157.7 3.35
30 S10 43.1 52.6 IL — 14 96 158.6 3.40
31 S11 44.7 42.1 CRS 0.5 14 192 137.1 3.01
32 S11 44.6 48.4 IL — 14 192 143.6 3.11

the table, CRS tests on the untreated soil were conducted us- tween initial total mass and final dry mass. For calculation of
ing rates in the 0.25%–1%/h range, while all tests on the ce- the volume of the hydrated cement a value of 2.15 was used
ment-treated soil employed the same rate of 0.5%/h. With for the specific gravity (Lea 2004). For the soil, the specific
one exception (test 10 in Table 1), the maximum pore pres- gravity was derived from the linear relationship existing be-
sure ratio (Duh/sv, where Duh is the excess pore pressure tween Gs and LOI for this particular soil (Santagata et al.
and sv is the vertical stress) measured in all the CRS tests 2008), based on measurements of the LOI conducted prior
performed on the untreated soil remained below the 30% to mixing the soil with cement. The void ratio was then cal-
threshold fixed by ASTM D4186 (ASTM 1989) (all tests culated as the ratio of the volume of water (at the end of the
were performed prior to the 2006 revision of the standard). curing stage all specimens were found to be essentially 100%
For all tests on the cement-treated soil the maximum pore saturated) and the sum of the volumes of the two solid
pressure ratio remained below 5%. phases. Figure 1 presents average volume phase diagrams for
Note that the values of the cement content provided in Table 1 pre-consolidation (end of curing) conditions for specimens
deviate somewhat from the target nominal values of 8%, prepared with 20%, 50%, and 100% cement by dry mass of
20%, 50%, and 100%. For simplicity the latter values (50% the soil. Note from Fig. 1, and from the data in Table 1, that
and 100%) are used in the remaining text. there is a trend of decreasing void ratio at the start of consol-
Calculation of the phase relations of the treated soil idation (i.e., at end of curing), with increasing cement con-
presents some challenges, in particular due to the need to es- tent. This trend is consistent with the cement and water
timate the specific gravity of the treated soil, which decreases added during mixing, the water consumed by the cement hy-
continuously during hydration (e.g., Lorenzo and Bergado dration reaction, and the deformations measured under the
2004). In this work, calculation of the void ratio of the curing surcharge. The deformation increased with decreasing
cement-treated organic soil was performed treating cement cement dosage.
and soil as separate phases, i.e., calculating the mass and vol- As shown in Table 1, each test has been assigned a se-
ume of each of these phases separately. With this assumption quential reference number. This is the same reference number
the masses of the two solid phases were calculated from the used in a previous publication (Santagata et al. 2008) focused
post-test oven-dried mass and the mix proportions used to on the behavior of the untreated soil. Tests 1–5 performed on
prepare each cement-treated soil specimen (accounting for intact block samples of the natural soil are omitted from
the amount of water reacting with the cement); while the Table 1, as they are not directly relevant to the research pre-
mass of the water phase was derived from the difference be- sented in this paper.

Published by NRC Research Press


1104 Can. Geotech. J. Vol. 48, 2011

Fig. 1. End of curing volume phase relations for 20%, 50%, and The elemental (C, H, N) composition of each of the frac-
100% cement-treated soil. Hydr., hydrated. tions was determined using a LECO CHN-2000 analyzer. In
this technique the carbon, nitrogen, and hydrogen content are
determined from analysis of the gases released by the com-
bustion of the soil sample.
The different fractions extracted from both the untreated
and cement-treated soil were further characterized employing
FTIR. This technique is especially suited for the characteriza-
tion of humic substances that have no regularly repeating
structural units, and are, instead, made of a variety of func-
tional groups. FTIR spectroscopy is based on the fact that
molecules vibrate at discrete energies in the infrared (IR) re-
gion of the electromagnetic spectrum. When the frequency of
the incident IR radiation matches the frequency of the partic-
ular vibrational mode (bending or stretching) of a molecule,
IR absorption occurs. FTIR analyses for this research were
conducted using a Perkin-Elmer 1600 spectrophotometer.
Chemical tests Each specimen, prepared in solid form using potassium bro-
pH measurements, fractionation tests, elemental analyses mide (KBr), was scanned 640 times from 370 to 4000 cm–1
using a LECO analyzer, and Fourier transform infrared spec- (the vibration of most organic molecules occurs in the mid-
troscopy (FTIR) analyses were conducted to complement the IR spectral region) with an 0.5 cm–1 interval and 2 cm–1 res-
consolidation experiments. olution.
pH measurements on the untreated soil were conducted us-
ing a Corning 44 pH meter, a pH glass electrode, and a calo- Results and discussion
mel (reference) electrode in accordance with ASTM D4972
(ASTM 2007). The same device was used for measuring the Overview of effects of cement on 1D consolidation
pH of the cement-treated soil. For these tests an amount of behavior
soil corresponding to a dry mass of 15 g was mixed with a Results from tests on soil with 0% (reconstituted samples)
cement slurry (with water to cement ratio of 0.5), and water and 20% PC (59 kg/m3) are presented in Fig. 2 to illustrate
was added to achieve an overall water to solids ratio of 4. the consistency in results between CRS and IL tests, and to
Measurements were conducted 1 h after mixing, as well as 1 highlight some of the changes in 1D consolidation behavior
and 7 days later. Between measurements the treated soil was induced by the addition of cement. As shown in Fig. 2a, the
kept in a tightly sealed glass container to prevent any mois- compression curves for 20% PC obtained from CRS and IL
ture loss. consolidation tests coincide, indicating reproducibility in the
Fractionation of both the untreated and treated soil was behavior of the cement-treated soil as well as no significant
performed to quantify the percentage of humin, humic acid, strain-rate effects (the two tests necessarily involve straining
and fulvic acid present (see section titled “Materials”). For of the soil at different rates). Similar observations apply to
this purpose, samples were prepared following the same pro- the untreated soil (see also Santagata et al. 2008). Figure 2a
cedure employed for the samples used for the consolidation is a plot of void ratio with effective stress, Fig. 2b includes
tests, except that no curing surcharge was applied. Instead, the change of permeability with void ratio, Fig. 2c shows
after compaction the samples were stored in a glass jar for data of coefficient of consolidation with effective stress, and
7 days. At this time a sample of mass corresponding to 15 g Fig. 2d of constrained modulus with effective stress. Figure 2a
of dry soil was fractionated employing the following proce- indicates that the most notable effect of treatment on the
dure: (i) the samples were treated with a 0.1 mol/L HCl sol- compression behavior is the increase in the preconsolidation
ution (with solution to solids ratio of 5) to remove any pressure, s p0 , which, based on determinations using the
carbonates so that the organic matter could be easily sepa- strain energy method (Becker et al. 1987), increases by
rated from the mineral portion of the soil; (ii) after continu- about 80% (from an average value of 50.2 kPa for the un-
ous shaking for 48 h the samples were centrifuged at treated reconstituted soil to approximately 94 kPa). As a re-
2000 revolutions/min for 15 min and the supernatant solution sult of the development of this cementation-induced s p0 , the
decanted; (iii) humic and fulvic acids were extracted using a compression curve of the treated soil is shifted to higher ef-
highly basic solution (pH > 12); (iv) the humic acid was then fective stress, with the treated soil able to carry almost
separated from the fulvic acid using a solution with pH < 1; twice the vertical effective stress sustained by the untreated
(v) the humic acids, which can be very strongly bonded to soil at the same void ratio.
mineral matter, were purified using an HF solution to elimi- Beyond s p0 the slope of the compression curve for the 20%
nate any additional inorganic matter. Note that this is the PC-treated soil is essentially constant, with values of the
same procedure used on the untreated soil (Santagata et al. compressibility index (Cc = 1.29–1.33) in the range of those
2008), except that it avoids pretreatment with a 0.1M HF sol- measured on the reconstituted untreated soil. The effects of
ution, as this step, which is used on the untreated soil to treatment are, instead, apparent in the swelling and reloading
eliminate silicates, would promote dissolution of some of the indices measured during unload–reload stages. In the over-
cementitious products. consolidation ratio (OCR) 1–2 range both the swelling and

Published by NRC Research Press


Bobet et al. 1105

Fig. 2. Effect of 20% cement addition on (a) compression behavior, (b) hydraulic conductivity, (c) coefficient of consolidation, and (d) con-
strained modulus of highly organic soil. mv, coefficient of volume compressibility.

the reloading indices decrease by more than 30% compared Fig. 3. Effect of 20% cement addition on excess pore pressure gen-
to the values measured for the untreated soil specimens. erated in CRS test.
Figure 2b shows that, at a given void ratio, the average hy- 200
Excess pore pressure/strain rate,

draulic conductivity of the treated soil is approximately 4.6


times greater than that of the reconstituted soil. Despite this
difference, the slope of the linear portion of the hydraulic 150
conductivity (k) plot (Ck = De/Dlogk) remains almost un-
Duh /ea (kPa/%/h)

changed with treatment. For the soil treated with 20% PC 0% PC


the average value of Ck/Cc is 0.66, in the range of the data 100 No. 13(CRS)
obtained for the untreated reconstituted soil (0.58–0.67). .
Note also the slight difference (∼15%–20%) between the
curves of hydraulic conductivity versus void ratio obtained 50 19% PC
from the IL and CRS consolidation tests. The effects of treat- No. 16(CRS)
ment on permeability are also highlighted in Fig. 3, which
shows the cumulative excess pore pressure generated during 0
loading and normalized by the measured strain rate for both 4.0 3.5 3.0 2.5 2.0 1.5
0% and 20% PC. It is seen that only 20% of the excess pore Void ratio, e
pressures produced in the reconstituted soil are generated in
the treated soil at the same void ratio. Given that the con-
strained modulus is not affected by treatment in the normally data) and untreated organic soil. Some discrepancy between
consolidated region (as discussed later), the results presented the IL and CRS data is observed, with the Cv values obtained
in Fig. 3 indicate an increase in the hydraulic conductivity of from the CRS test being consistently two times higher than
the soil with the addition of cement. the IL data. More importantly, the results show that the coef-
Figure 2c shows the variation of the coefficient of consol- ficient of consolidation increases with the addition of PC by
idation with stress level for both the treated (IL and CRS approximately one order of magnitude over the stress range

Published by NRC Research Press


1106 Can. Geotech. J. Vol. 48, 2011

investigated. This is a reflection of the increase in k with the concentration reduces the double-layer thickness of the phyl-
treatment noted above. losilicate minerals present in the organic soil, promoting floc-
The values of the constrained modulus of the treated soil culation and aggregation of particles. Second, the formation of
are plotted against vertical effective stress in Fig. 2d. In the the C–S–H and C–A–H gels binds soil particles together, fur-
overconsolidated region, the treated soil exhibits generally ther contributing to aggregation and flocculation (Tremblay et
stiffer response with higher constrained modulus, but once al. 2001; Al-Rawas 2002) of the soil particles. As a result of
consolidation proceeds past the preconsolidation pressure, these influences a more open fabric is formed, with macro-
the constrained modulus of the treated soil coincides with pores, which serve as the main channels for flow, having size
that of the untreated reconstituted soil. that increases with increasing cement content. These hypothe-
ses are supported by previous observations of the microstruc-
Effect of cement dosage on 1D consolidation behavior and ture of cement-treated clay conducted by Chew et al. (2004)
structure of cement-treated soil using scanning electron microscopy (SEM) and mercury in-
Having highlighted the general effects that the addition of trusion porosimetry. SEM observations show that the addition
cement has on the 1D compression behavior, this section ad- of cement produces a flocculated structure, and both analyses
dresses the role played by the cement dosage. Specifically, indicate an increase in the size of the openings formed
Fig. 4 presents the data for CRS tests conducted with cement amongst soil particle clusters with increasing cement content.
dosage varying between 0% (untreated reconstituted soil) and Particle-size analyses on untreated and cement-treated soil
approximately 100% (cement factor, 296 kg/m3). Similar to conducted by both Rao and Rajasekaran (1996) and Chew et
Fig. 2, Fig. 4a plots void ratio with effective stress, Fig. 4b al. (2004) also demonstrate an increase in particle size with
void ratio with permeability, Fig. 4c coefficient of consolida- cement content.
tion with effective stress, and Fig. 4d constrained modulus Note that it is not expected that the same trend of increas-
with effective stress. While, as discussed for 0% and 20% ing hydraulic conductivity with cement content would extend
PC, the results of the CRS and IL tests are found to be con- to the larger cement dosages used in cutoff walls. In such
sistent, for clarity, the data for the latter are not included in cases the high cement dosages employed would ultimately
this figure. Key results are summarized in Table 2. cause a decrease in kv.
The compression curves of the reconstituted and PC-treated Values of Ck and Ck/Cc derived from the data presented in
soil are shown in Fig. 4a. The figure shows that the greater Fig. 4 are summarized in Table 2. It is observed that for ce-
the cement dosage, the greater the cementation-induced pre- ment contents up to 50% there is no clear trend between ce-
consolidation stress, the more extended the reloading region, ment content and either Ck or Ck/Cc. In fact, for both these
and the greater the vertical effective stress that can be sus- parameters the data for the cement-treated soil fall within the
tained at a given void ratio. range observed for the untreated reconstituted soil. However,
The figure also highlights the trend of decreasing value of with 100% cement, the values of both Ck and Ck/Cc fall
the void ratio at the start of consolidation (i.e., at end of cur- clearly below the range measured on all other specimens.
ing; see data in Table 1) with increasing cement content. This The decrease in Ck/Cc is especially of interest, as it reflects
trend is consistent with the cement and water added during the fact that a smaller percentage of the voids that contribute
mixing, the water consumed by the cement hydration reac- to the compressibility of the soil contribute to its hydraulic
tion, and the deformations measured under the curing sur-
conductivity. This suggests that the fabric formed as a result
charge.
of treatment with 100% cement involves some internal – not
While the preconsolidation stress increases with treatment, connected pore space. Similar observations are reported by
the compression index in the normally consolidated region
Kang (2011) for a cement-treated inorganic clay.
does not show significant changes with cement addition. As
The coefficient of consolidation is plotted against vertical
shown in Table 2, there is no clear trend with cement dosage
effective stress in Fig. 4c. In general, at the same vertical ef-
and, regardless of the cement content, the values of Cc fall
fective stress, Cv increases with treatment. Similar to pre-
within the range observed for the untreated reconstituted soil.
vious observations regarding s v0 and kv, the effects of
Figure 4b shows the increase in hydraulic conductivity of
the organic soil with treatment. Although the initial void ratio treatment with 8% PC are small. The increase in Cv is due
decreases gradually with increasing cement content, the hy- to the increase in both hydraulic conductivity and constrained
draulic conductivity of the treated soil increases with cement modulus (discussed later). Similar results have been obtained
percentage at any given void ratio. With 8% PC the effect of for other soils treated with lime and cement (Broms 1999;
the treatment is modest, with a 18%–26% increase in k at the Cortellazzo and Cola 1999; Kassim and Clarke 1999; Kang
same void ratio. However, a much greater increase in hy- 2011).
draulic conductivity is observed with higher PC contents, The effect of cement treatment on the constrained modulus
and for 100% PC the hydraulic conductivity is over one order is illustrated in Fig. 4d, which shows that while cement treat-
of magnitude greater than that of the untreated soil. This ob- ment leads to an increase in the constrained modulus in the
servation is in agreement with past findings (Broms and Bo- overconsolidated region, the relationship between modulus
man 1979; Brandl 1981; Buensuceso 1990; Townsend and and effective stress is independent of cement content in the
Klym 1996; Cortellazzo and Cola 1999; Kang 2011). The in- normally consolidated region.
crease in hydraulic conductivity with treatment is caused by The average increase in preconsolidation pressure with
the change in the fabric of the soil as a result of the cement treatment, shown in Fig. 4a, is summarized in Fig. 5, which
reactions. First, the released Ca2+ and OH– increase the pH shows that the increase in s p0 is relatively modest for low val-
and the ionic strength of the solution; the increase in the Ca2+ ues of cement (50% and 88% increase in s p0 with 8% and

Published by NRC Research Press


Bobet et al. 1107

Fig. 4. Effect of cement treatment (8%–100% PC) on: (a) compression behavior, (b) hydraulic conductivity, (c) coefficient of consolidation,
and (d) constrained modulus of highly organic soil.

Table 2. Summary of results on cement treated soil.


PC (%) Cc, 2  4s P0 Cr, OCR 4–2 Ck Ck/Cc Ca Ca/Cc
0 1.548 ± 0.207 0.266 ± 0.055 0.95 ± 0.09 0.60 ± 0.04 0.123 0.095
8 1.444 0.057 0.89 0.64 0.106 0.085
20 1.326 0.026 0.87 0.66 0.080 0.063
50 1.699 0.013 0.91 0.61 0.070 0.042
100 1.382 0.005 0.73 0.54 0.029 0.021
Note: Ca, creep coefficient.

20% PC, respectively), but much sharper at higher dosages, With this normalization the compression curves of recon-
as s p0 increases by approximately 800% and 2300% with stituted–remolded soils all fall around a unique line termed
50% and 100% PC, respectively. the intrinsic compression line (ICL), and the in situ state for
Further comparisons between the reconstituted and treated a variety of sedimentary natural clays in the normally con-
soils can be made within the framework proposed by Burland solidated state plots in a band to the right of the ICL. The
(1990), which relies on the normalization of the compression best-fit curve through these data is termed the sedimentation
curves through the use of the void index, Iv. Iv is defined compression line (SCL). At the same void ratio the SCL lies
based on e100 and e1000 , the intrinsic void ratios, i.e., the at vertical effective stresses approximately five times higher
void ratios of the soil in the reconstituted–remolded state at than the ICL, implying that most naturally sedimented clays
stresses of 100 and 1000 kPa, respectively, as follows: can support five times higher vertical effective stress than re-
constituted clays. Burland (1990) as well as others have
e  e100
½1 Iv ¼  shown that the in situ state of some clays can fall substan-
e100  e1000 tially above the SCL.

Published by NRC Research Press


1108 Can. Geotech. J. Vol. 48, 2011

Fig. 5. Effect of cement treatment (8%–100% PC) on preconsolida- level in the virgin compression region as the sum of two
tion stress. terms:
 0 b
 
sp
½2 e ¼ e þ De ¼ e þ Dei 0
sv
where e* is the void ratio of the reconstituted soil at the same
vertical effective stress; De is the difference between the void
ratio of the structured soil and that of the reconstituted soil at
a given vertical effective stress s v0 ; Dei is the value of De de-
termined in correspondence to the preconsolidation stress s p0 ;
and b, which is termed the compression destructuring index,
is a parameter that quantifies the rate of reduction in De with
stress level. Data presented by Liu and Carter (2000) suggest
that for natural soft clays, b generally falls in the 0.3–1 range
(although higher values of b are reported). Much smaller va-
lues of b (<0.1) are obtained from the first loading stage (i.e.,
prior to the unload–reload stage) of the 1D compression tests
While Burland’s (1990) framework was developed for nat- on the 50% and 100% cement soil specimens (see Fig. 6).
ural soils, Kang and Santagata (2006) suggest that the intrin- This indicates limited damage to the cement-induced struc-
sic values of the untreated reconstituted soil may be used to ture produced by 1D loading.
normalize the data of cement-treated soil, and thus the frame- The compression behavior of the soil following the unload-
work can provide insight into the behavior and the degree of ing stage provides additional insight. As shown in Figs. 4
structuring of cement-treated soils. Figure 6a, which presents and 6a, and in more detail in Fig. 6b for 50% cement, it is
the compression curves of all cement-treated specimens in found that following the unload–reload cycle, the compres-
the void index (Iv)–logs p0 space, illustrate the results of such sion curve for the 50% and 100% PC specimens “overshoots”
the original virgin compression line. This appears to be the
a normalization. Also included in this figure are the data for result of the additional cement-induced structure developed
the reconstituted soil, which fall on the ICL, and the com- during the creep stage that precedes unloading. Estimates of
pression curves for the intact natural soil based on tests on the destructuring index b for the second loading stage are in
block samples (Santagata et al. 2008). It is seen that the nor- the 0.4–0.6 range for the 50% and 100% PC specimens, indi-
malized compression curves of the cement-treated soil are ap- cating a more rapid destructuring of this component of the
proximately parallel to the intrinsic and sedimentation soil structure; with continued loading the compression curve
compression lines, and that as the cement dosage increases rejoins that determined from the first loading stage. Similar
the curves shift to the right, reflecting, as discussed earlier, observations apply also to the data obtained following differ-
an increased cement-induced structure. While the curves for ent curing procedures, which is discussed in the section titled
8% and 20% cement fall between the ICL and the SCL, the “Effects of curing duration and curing surcharge”.
curves for 50% and 100% cement fall very close to the SCL, Although the compression index does not show significant
suggesting a degree of structuring similar to that typical of change with treatment, the swelling index and the recompres-
many natural sedimentary clays. Similar observations were sion index decreased with treatment. For example, as shown
reported by Kang and Santagata (2006) for a cement-treated in Table 2, which summarizes data from the unload–reload
inorganic clay. loops over the same OCR range, the recompression index de-
Figure 6a also provides insight into the nature of the struc- creased by a factor of approximately 2, 5, 10, and 25 follow-
ture generated in the cement-treated clay. Burland (1990) ing treatment with 8%, 20%, 50%, and 100% PC, respectively.
shows that for natural clays whose stress state lies close or Finally, creep measurements were performed on soil
above the SCL, the post-yield compression curve is markedly specimens treated with different dosages of cement, all at the
steeper than the SCL, and that at high stresses it tends to ap- maximum vertical stress level reached in the IL tests
proach the ICL as a result of the progressive collapse of the (∼1600 kPa). As shown in Fig. 7, these data show that in
soil structure. Baudet and Stallebrass (2004) refer to the struc- the secondary compression regime the log time–deformation
ture associated with this type of behavior as a “meta-stable” curve is linear (constant creep coefficient Ca) for more than
structure. As shown in Fig. 6a, and discussed in more de- two log cycles of secondary compression, with Ca decreasing
tail by Santagata et al. (2008), this is the behavior observed with increasing cement content (by over a factor of 4 for
when testing the intact block samples. In contrast, Fig. 6a 100% PC; see also Table 2). As indicated in Fig. 7, Ca of
shows that upon first loading, the post-yield compression the reconstituted untreated soil is 0.122, while for 8%, 20%,
curves of the 50% and 100% cement-treated soil clay re- 50%, and 100% PC the values are 0.108, 0.080, 0.064, and
main practically parallel to the SCL. This is evidence of 0.029, respectively, all measured at the same stress level.
the stable nature of the structure generated as a result of Given that Ca is known to vary significantly, depending on
cement treatment. stress level and OCR, it is more insightful to examine the var-
Post-yield structure degradation can be described using the iation of Ca/Cc with cement addition. For a given soil this ra-
model proposed by Liu and Carter (1999, 2000), which ex- tio is generally observed to remain constant, reflecting the
presses the void ratio of structured soil at any given stress similar nature of the mechanisms that contribute to primary

Published by NRC Research Press


Bobet et al. 1109

Fig. 6. Compression curves of cement-treated soil plotted using void index: (a) 0%–100% cement; (b) detail of different testing stages for 50%
curve.

and secondary consolidation. Moreover, for a given soil loaded soils, and that it increased with increasing curing
type Ca/Cc is known to vary within a fairly limited range time. Åhnberg et al. (2001) observed that the unconfined
(Mesri and Castro 1987). As shown in Table 2 and Fig. 8, compressive strength of cement-treated peat increased when
Ca/Cc decreases markedly with cement addition, from val- the surcharge was applied immediately after treatment.
ues at the high end of the range typical of peats and mus- The beneficial effects of prolonged curing are likely to be
kegs, to values characteristic of inorganic clays (50% PC), twofold. First, the increase in the duration of secondary com-
and finally to values generally measured on granular soils pression will lead to an increase in the aging-induced precon-
(100% PC). This variation in Ca/Cc reflects the modifica- solidation stress. Second, prolonged curing will allow an
tions in the nature of the “new” geomaterial formed as a increased time for the cement hydration reactions to take
result of treatment with different percentages of cement. place. In this research, the effects of curing time were eval-
Given the small change in the compression index with ce- uated by conducting experiments on the organic soil treated
ment addition noted above, the decrease in Ca/Cc is mainly with 50% and 100% cement cured for 14 and 28 days. These
due to the reduction of the secondary compression index cement contents were selected because they yielded the most
with cement content. significant changes in engineering properties, and because, as
discussed in the section titled “pH conditions promoting for-
Effects of curing duration and curing surcharge mation of calcium silicate gel”, the pH values measured with
The importance of the curing process on the properties of these cement contents indicated conditions that would pro-
chemically stabilized soils has been highlighted by many re- mote continued hardening of the calcium silicate gel. The re-
searchers. Hwang et al. (2004) observed that stiffness and un- sults obtained from testing these samples indicate no effect of
confined compressive strength of organic soils treated with curing period on the compression behavior of the 50% ce-
PC and lime increased with increasing curing time and cur- ment soil samples. For 100% cement the effects are small,
ing surcharge. Hebib and Farrell (2003) observed that the un- with a slight increase in the preconsolidation stress (∼7%)
confined strength of PC-stabilized peat increased when the and in the coefficient of consolidation (<10%) measured on
soil was preloaded during curing. They also observed that the specimens cured for 28 days, relative to the 14 day re-
the unconfined compressive strength was higher for pre- sults. While small, these improvements are likely related to

Published by NRC Research Press


1110 Can. Geotech. J. Vol. 48, 2011

Fig. 7. Effect of cement treatment (8%–100% PC) on creep coeffi- Fig. 9. Effect of surcharge stress on compression behavior of cement
cient. treated soil.

48, 96, and 192 kPa, which represent approximately 2.5, 5,


and 10 m thick fills with a unit weight of 19 kN/m3. Figure 9
presents the compression curves for six tests (two for each
of the stress levels investigated), some of which also in-
volved unload–reload stages. Focusing on the first loading
stage, it is seen that while the specimens cured under a
higher surcharge have a lower initial void ratio, all the com-
pression curves reach the same virgin compression line.
This is consistent with previous observations by Tremblay
et al. (2001) that the virgin compression line of cement-
treated soil is independent of the initial curing void ratio.
The average preconsolidation stress increases from 403.4 ±
1.7 kPa (surcharge = 48 kPa) to 471.4 ± 5.0 kPa (96 kPa),
i.e., an increase of 68 kPa for a surcharge increase of
48 kPa, to 672.7 ± 24.4 kPa (192 kPa), i.e., an increase of
269.3 kPa for a surcharge increase of 144 kPa. The im-
provement in the preconsolidation stress is mainly due to
the decrease in the end of curing void ratio that dictates at
which stress level upon reloading in the consolidation cell
the virgin compression line is reached. This implies that
Fig. 8. Effect of cement treatment (8%–100% PC) on Ca/Cc. the yield stress of the treated soil can be improved by cur-
ing the soil under a higher surcharge for the same duration.
Analysis of the data for the three different curing stresses
also shows that the constrained modulus and hydraulic con-
ductivity increase with increasing curing surcharge in the
overconsolidated region, while the effects of surcharge are
negligible in the normally consolidated region. As a result,
the coefficient of consolidation also increases with surcharge,
with the increase in the constrained modulus as the main fac-
tor.

Chemistry of cement-treated soil


The pH measurements, fractionation tests, LECO elemen-
tal analyses (C, H, N), and FTIR analyses were conducted to
complement the consolidation experiments. The overall goal
the continuous hardening of the C–S–H gel during the time of this portion of the experimental work was to gain insight
frame investigated. Overall, the small changes in behavior into the interaction between the organic matter and the ce-
observed with increased curing time justify the selection of ment, and in particular to assess the stability of the organic
the 14 day curing period used for the majority of the tests. matter following treatment, evaluate the reactivity of the dif-
The effect of the curing surcharge was investigated exclu- ferent functional groups present, and investigate the composi-
sively for soil treated with 50% cement using surcharges of tional changes of the different fractions of the treated soil.

Published by NRC Research Press


Bobet et al. 1111

pH conditions promoting formation of calcium silicate gel Fig. 10. Effect of cement addition on soil pH.
The increase in the preconsolidation pressure and the shift 14
of the compression curves to the higher vertical effective 12.6
stress described are indications of the development of a stiffer 12
soil structure with the treatment. The structure is developed
mainly as a result of the formation of the cement hydration 10

pH
products and their reactions with the soil particles. All these
reactions are very sensitive to the chemical environment, es- 8
1 day after mixing
pecially the pH of the soil–cement system. As soon as ce-
7 days after mixing
ment gains access to water, formation of what are sometimes 6
referred to in the soil cement literature (e.g., Bergado et al.
1996) as “primary cementitious products” starts. These in- 4
clude calcium silicate hydrate (C–S–H), calcium aluminate 0 20 40 60 80 100 120
hydrate (C–A–H), and calcium hydroxide (Ca(OH)2). Cal- Cement content (%)
cium silicate gel (C–S–H), which is formed from reaction of
Table 3. Results of fractionation tests on cement treated soil.
the cement’s tricalcium silicate (C3S) and dicalcium silicate
(C2S) with water, is the major source of the strength of the Cement Nonextracted fraction Humic Fulvic
soil–cement mixture. Its formation requires a pH greater than content (%) (%) acid (%) acid (%)
12.6. Following these reactions, Ca2+ and OH– ions are re- 0 73.86 23.71 2.43
leased into the pore water. The released OH– ions increase 20 99.18 0.59 0.23
the pH of the pore water. Under high pH conditions, silica 50 99.93 0.01 0.06
(SiO2) and alumina (Al2O3) from the clay minerals are dis- 100 99.99 0.00 0.01
solved and combine with the released Ca2+ and OH– ions to
form “secondary” cementitious products (C–S–H and C–A–H).
The amount of silica and alumina dissolved, and hence the Table 3 represent the average values from these two measure-
amount of secondary cementitious products formed as a re- ments. The results show that the amount of humic and fulvic
sult of the pozzolanic reaction, are markedly dependent on acid fractions extracted decrease substantially with increasing
the pH of the soil–cement system. cement content. For example, less than 3% of the humic acid
Figure 10 shows the pH values measured with different of the untreated soil could be extracted from the soil treated
percentages of cement 1 h and 1 day after mixing. The un- with 20% cement. With 50% cement this percentage de-
treated soil (0% PC) was slightly acidic with a pH value of creases to less than 0.1%, and with 100% cement the amount
5.98. Following treatment with 8% PC, the pH increased rap- of humic acid extracted was too small to be measured accu-
idly to 11.13 1 h after treatment. No further change in pH rately. The fulvic acid extracted also decreased with increas-
was observed with additional time. With 20% PC the pH in- ing cement, but to a lesser degree: with 20% PC, only 10% of
creased to 11.98 in 1 h, and continued to increase to 12.20 the fulvic acid was extracted, and this amount decreased to
after 1 day, with no subsequent increase. With 50% PC, the 3% and 0.5% with 50% and 100% cement, respectively.
pH increased to 12.20 in 1 h, and to 12.61 after 1 day. With Given that the addition of cement is associated with an in-
100% PC, the pH increased to 12.32 in 1 h and to 12.77 after crease in the pH (see Fig. 10), and given that under highly
1 day. In each case, no further pH increase with time was ob- alkaline conditions, dissolution of humic and fulvic acids is
served. known to occur (see section titled “Materials”), the question
These measurements are in good agreement with the pre- may arise as to the effect of pH on the stability of the organic
consolidation stress data (Fig. 5), and the sharp increase in matter. Loss on ignition tests conducted at the end of the
s p0 with 50% and 100% PC can be related to the increase in consolidation tests and elemental analyses of the three frac-
pH to values above 12.6, which allows formation and contin- tions extracted from the cement-treated samples demonstrate
ued hardening of the primary cementitious products. The pH that no dissolution of the organic matter occurred during the
values measured also suggest that for the soil–cement mix- curing process, and that organic matter was not lost during
tures examined in this research, the contribution to the extraction and fractionation. Thus, the significant reduction
strength from the “secondary” cementitious products was of the extracted humic and fulvic acids following treatment
negligible, as for values of pH < 14, no major dissolution of with cement indicates that the organic matter was bound or
silica and alumina occurs (S. Diamond, personal communica- at least encapsulated by the cement hydration products and
tion, 2010). was present in the mineral fraction of the soil.
As done for the soil alone, following fractionation, the hu-
Interaction of organic matter and cement mic substances were further characterized employing elemen-
As discussed in the section titled “Materials”, the organic tal analysis and FTIR. Table 4 summarizes the masses of
portion of the soil used in this experimental program is com- carbon, nitrogen, and hydrogen present in each of the three
prised of humic substances that can be classified according to fractions (for cement contents greater than 20% no analysis
their solubility as humin, humic acid, and fulvic acid. Table 3 could be conducted on the humic and fulvic acids due to the
summarizes the results of the fractionation tests, including insufficient amounts extracted). All data pertain to samples
the data obtained for the untreated soil. As mentioned earlier, with the same amount of dry soil (15 g). The data presented
two independent measurements were conducted for each ce- in Table 4 illustrate that the C, N, and H extracted as humic
ment content on two separate samples. The data presented in acids from the 20% cement sample decreased significantly

Published by NRC Research Press


1112 Can. Geotech. J. Vol. 48, 2011

Table 4. Mass of C, N, and H present in fractions extracted from untreated and cement treated soil.
C (g) H (g) N (g)
Cement Nonextracted Humic Fulvic Nonextracted Humic Fulvic Nonextracted Humic Fulvic
content (%) fraction acid acid fraction acid acid fraction acid acid
0 1.601 1.826 0.173 0.354 0.172 0.011 0.114 0.132 0.005
20 2.981 0.048 0.014 0.482 0.006 0.002 0.196 0.006 0.002
20 3.106 0.049 0.017 0.500 0.006 0.003 0.196 0.006 0.002
50 3.126 — — 0.540 — — 0.180 — —
50 3.259 — — 0.562 — — 0.202 — —
100 3.210 — — 0.630 — — 0.150 — —
100 3.540 — — 0.660 — — 0.210 — —

compared to the natural soil. In contrast, the carbon, nitrogen, carbonate (CaCO3) from the reaction of CO2 with the
and hydrogen contents of the nonextracted fraction increased Ca(OH)2 formed by cement hydration. The greater the ce-
with cement content. Again this is evidence that a significant ment percentage, the greater the amount of Ca(OH)2 avail-
portion of the organic matter present in the natural soil is able; hence, the greater the amount of CaCO3 formed by
bound–encapsulated by the cement hydration products. carbonation, as reflected in the FTIR peaks. However, car-
Additional perspectives are obtained by examining the bonation is an inevitable reaction following exposure to air,
composition of the fulvic and humic acids extracted from the and therefore has no specific significance for this study. The
natural soil and the sample with 20% cement, as done in band in the 1576–1772 cm–1 region increases gradually
Table 5, in which the carbon, nitrogen, and hydrogen con- with respect to the baseline in the 2000–2500 cm–1 region.
tents are expressed as a percentage of the mass of humic or In the mineral fraction this is the only band that represents
fulvic acid extracted. These data indicate that the chemical the organic functional groups, which include –C=O stretching
composition of the humic and fulvic acids extracted from the of –COOH (1720 cm–1); C=O stretching of amide groups
soil treated with 20% PC changed compared to that of the (1630–1660 cm–1); aromatic C=C stretching and (or) asymmet-
acids extracted from the natural soil. Specifically, with addi- ric –COO– stretching (1610 cm–1); as well as COO– symmet-
tion of cement the organic carbon percentage decreased, ric stretching, N–H deformation, and +C=N stretching
while the hydrogen and nitrogen percentage increased. This (1590–1517 cm–1). All the organic functional groups listed
suggests that different functional groups have different reac- here are composed mainly of C, H, and N. The increase in
tivity with the cement. the C, H, and N contents in the nonextracted PC-treated
Further insight about the reactivity with cement of func- soil can be interpreted as the retention in the mineral frac-
tional groups of the organic matter is provided by the FTIR tion of these functional groups by cementation.
spectroscopy analyses, which were performed on each frac- Figure 11b presents the FTIR spectra of the humic acid
tion of both the natural and the cement-treated soil, except fraction extracted from the untreated and treated soil. The rel-
for the fulvic and humic acids extracted from the 100% PC atively high intensity band in the 3300–3400 cm–1 region of
treated soil due to the small amounts available. The objec- the untreated soil indicates the presence of phenolic OH (OH
tives of the FTIR analysis were to investigate the composi- stretch of phenolic OH). The increase in the intensity of the
tional changes of the different fractions of the treated soil broad band in the 3300–3600 cm–1 region may indicate that
and evaluate the reactivity of the different functional groups part of the Ca(OH)2 and the C–S–H gel were dissolved and
present. Two analyses were performed on each fraction for extracted under the highly acidic environment that was cre-
consistency. These data are reported in Figs. 11a–11c. ated during fractionation to precipitate humic acid (pH = 1).
Although the frequencies corresponding to the various vibra- The intensity of the 1234–1262 cm–1 region, which is where
tion modes (symmetric or asymmetric stretching and bending the –C–OH stretching band of phenolic OH occurs, did not
vibration modes) of organic functional groups have been change with treatment. This is an indication that phenolic
studied extensively, the similarity of the chemical composi- OH does not react with cement, which is in agreement with
tion of organic functional groups makes it particularly diffi- findings from other researchers (Pollard et al. 1991). The
cult to accurately quantify the sources responsible for small peak in the 1718 cm–1 region (corresponding to the
changes in the FTIR spectra; a qualitative analysis however C=O stretching band of –COOH) present in the untreated
provides sufficient insight into the chemical reactions be- soil disappeared in the humic acid fraction of the treated
tween the soil and the cement. soil. At the same time the intensities of the bands at
Figure 11a shows the FTIR spectra of the nonextracted 1400 cm–1 and in the 1546–1600 cm–1 region (which reflect
fraction of the treated soil (mineral fraction). The most sig- the symmetric and antisymmetric stretching of COO–) in-
nificant change in the spectrum with cement treatment is the creased. These results do not necessarily indicate the reaction
gradual increase in the intensity of the broad band in the of the carboxyl group (COOH) with cement, but rather the
3000–3600 cm–1 region. This region represents the symmet- change of the carboxyl group to carboxylate group (COO–)
ric and asymmetric stretching of O–H. According to Yousuf under the highly acidic conditions used for precipitation of
et al. (1995), the increase in intensity of this band indicates the humic acid during fractionation.
the formation of Ca(OH)2 and of the C–S–H gel. The sharp Figure 11c presents the FTIR spectra of the fulvic acid. The
increase in intensity of the bands in the 1415 and 872 cm–1 strong band around 3390 cm–1 and the small band in the 1245–
regions is due to carbonation, i.e., the formation of calcium 1251 cm–1 region in the untreated soil indicate the presence of

Published by NRC Research Press


Bobet et al. 1113

Table 5. Percentage of C, N, and H present in humic and fulvic acid fractions extracted from 0% and
20% PC soil.

Humic acid Fulvic acid


Cement content (%) C (%) H (%) N (%) C (%) H (%) N (%)
0 49.9 4.7 3.6 41.2 2.5 1.1
20 45.4 5.6 5.2 34.8 5.6 4.7
20 45.4 5.5 5.2 39.9 6.3 5.5

Fig. 11. Effect of cement addition on FTIR spectra of (a) mineral gradually with treatment, while the intensities in the 1560
fraction, (b) humic acid, and (c) fulvic acid. and 1630 cm–1 region increased. These changes are the op-
posite of the results observed in the humic acid fraction,
which indicate the change of the carboxylate group
(COO–) to the carboxyl group (COOH) under the highly
basic solution created during fractionation.

Summary and conclusions


The 1D consolidation behavior of a highly organic soil
treated with Portland cement was investigated by performing
constant rate of strain (CRS) and incremental loading (IL)
consolidation tests on specimens obtained from soil–cement
samples cured under a surcharge of 48 kPa for 14 days. The
cement dosage varied from 8% to 100% by dry mass of the
soil, corresponding to values of the cement factor in the 24–
296 kg/m3 range. The most evident effect of treatment with
cement is the increase in the soil’s preconsolidation stress.
The increase is fairly modest for low dosages of cement, but
as large as eightfold and 25-fold for cement dosages of 50%
(148 kg/m3) and 100% (296 kg/m3), respectively. As a result
of the development of this cementation-induced preconsolida-
tion stress, the compression curves shift to higher effective
stress levels: the higher the cement content the greater the ef-
fective stresses that the soil can sustain at the same void ra-
tio. Additionally, the compressibility of the cement-treated
soil measured in the overconsoliated region also shows a sig-
nificant decrease with increasing cement. These two effects
are an indication of the effectiveness of the treatment in re-
ducing settlements.
In contrast to the above, the compression index (Cc) in the
normally consolidated region does not show any significant
change with cement addition. However, the structure devel-
oped as a result of the addition of cement appears stable,
and does not show significant degradation as a result of 1D
loading beyond the preconsolidation stress.
The hydraulic conductivity and coefficient of consolidation
of the treated soil increase with the addition of cement. The
increase in hydraulic conductivity is caused by the change in
the soil fabric, in particular the increase in size of the macro-
pores serving as the main channels for flow, which arises
from chemical reactions with the cement. Associated with
the increase in hydraulic conductivity is the increase in coeffi-
cient of consolidation: compared with the untreated reconsti-
tuted soil, Cv increases by 1.4 times with 8% PC (24 kg/m3),
8 times with 20% (59 kg/m3) PC, and about 38 times with
phenolic OH. Similarly, the small band at 2936 cm–1 corre- 50% (148 kg/m3) and 100% (296 kg/m3) PC.
sponds to the C–H stretching of the phenolic –CH2–. The fact The effects of treatment on the creep behavior of the or-
that these regions did not change with treatment indicates ganic soil were investigated by performing long-term sus-
the low reactivity of phenolic OH and –CH2– with cement. tained loading tests at the maximum effective stress reached
The sharp peak at 1400 cm–1 in the untreated soil decreased in the consolidation tests (∼1600 kPa). The creep coefficient,

Published by NRC Research Press


1114 Can. Geotech. J. Vol. 48, 2011

Ca, measured at this stress level is shown to decrease with ASTM standard D2487. American Society for Testing and
cement content, by as much as a factor of 4 with 100% PC Materials, West Conshohocken, Pa.
(296 kg/m3). As a result Ca/Cc also decreases. With 50% ASTM. 2007. Standard test method for pH of soils. ASTM standard
(148 kg/m3) PC, this ratio goes from the value of 0.095 D4972. American Society for Testing and Materials, West
measured on the untreated reconstituted soil to a value of Conshohocken, Pa.
0.04 typical of inorganic clays. With 100% PC (296 kg/m3), Baudet, B., and Stallebrass, S. 2004. A constitutive model for
Ca/Cc further decreases to 0.024, falling in the range of gran- structured clays. Géotechnique, 54(4): 269–278. doi:10.1680/geot.
2004.54.4.269.
ular soils (0.02 ± 0.01).
Becker, D.E., Crooks, J.H.A., Been, K., and Jefferies, M.G. 1987.
Fractionation tests and chemical and FTIR analyses pro-
Work as a criterion for determining in situ and yield stresses in
vide insight into the interaction between soil organic matter clays. Canadian Geotechnical Journal, 24(4): 549–564. doi:10.
and cement. The fractionation tests show that with increasing 1139/t87-070.
cement dosage, decreasing amounts of humic acid and fulvic Bergado, D.T., Anderson, L.R., Miura, N., and Balasubramaniam, A.
acid can be extracted from the soil–cement mixture. This in- S. 1996. Improvement of soft ground. American Society of Civil
dicates that a significant portion of the organic matter is Engineers (ASCE) Press, New York.
bound–encapsulated in the cement matrix. Moreover, elemen- Brandl, H. 1981. Alteration of soil parameters by stabilisation with
tal analyses show a change in the composition of the humic lime. In Proceedings of the 10th International Conference on Soil
acids and fulvic acids extracted from the cement-treated soil. Mechanics and Foundation Engineering, Stockholm, Sweden, 15–
This suggests that not all functional groups have the same re- 19 June 1981. A.A. Balkema, Rotterdam, the Netherlands.
activity with the cement. This hypothesis is supported by the pp. 587–594.
FTIR analyses, which show, for example, that phenolic Broms, B.B. 1999. Design of lime, lime/cement and cement columns.
groups have limited reactivity with the cement. In Dry Mix Methods for Deep Soil Stabilization, Proceedings of
While the research has focused on a particular soil, the ex- the International Conference on Dry Mix Methods for Deep Soil
perimental approach and the fundamental conclusions on the Stabilization, Stockholm, Sweden, 13–15 October 1999. A.A.
consolidation behavior of cement-treated organic soil and on Balkema, Rotterdam, the Netherlands. pp. 125–153.
the interaction between organic matter and cement can be ex- Broms, B., and Boman, P. 1979. Stabilisation of soil with lime
columns. Ground Engineering, 12(4): 23–32.
trapolated to other organic soils of a similar nature.
Bruce, D.A., and Bruce, E.C. 2003. The practitioner’s guide to deep
Acknowledgments mixing. In Grouting and Ground Treatment, Proceedings of the
Third International Conference, New Orleans, La., 10–12
This work was supported by the Joint Transportation Research February 2003. Geotechnical Special Publication No. 120. Edited
Program administered by the Indiana Department of Transporta- by L.F. Johnsen, D.A. Bruce, and M.J. Byle. American Society of
tion and Purdue University under project no. C-36-50U, grant Civil Engineers, Reston, Va. pp. 474–488.
SPR-2460. The contents of this paper reflect the views of the au- Buensuceso, B.R. 1990. Engineering behavior of lime treated soft
thors, who are responsible for the facts and the accuracy of the Bangkok Clay. Ph.D. thesis, Asian Institute of Technology,
data presented herein. The contents do not necessarily reflect the Bangkok.
official views or policies of the Federal Highway Administration Burland, J.B. 1990. On the compressibility and shear strength of
and the Indiana Department of Transportation, nor do the con- natural clays. Géotechnique, 40(3): 329–378. doi:10.1680/geot.
tents constitute a standard, specification or regulation. The writers 1990.40.3.329.
wish to acknowledge the help of Dr. G.S. Premachandra of Pur- Butcher, A.P. 2005. The durability of deep wet mixed columns in an
organic soil. In Proceedings of the International Conference of
due University’s Agronomy Department for assistance with the
Deep Mixing Best Practices and Recent Advances, Deep Mixing
fractionation and FTIR spectroscopy.
’05, Stockholm, Sweden, 23–25 May 2005. Swedish Deep
Stabilisation Research Centre, Linköping, Sweden. Vol. 1,
References
pp. 47–54.
Åhnberg, H., and Holm, G. 1999. Stabilization of some Swedish Chew, S.H., Kamruzzaman, A.H.M., and Lee, F.H. 2004. Physico-
organic soils with different types of binder. In Dry Mix Methods chemical and engineering behavior of cement treated clays. Journal
for Deep Soil Stabilization, Proceedings of the International of Geotechnical and Geoenvironmental Engineering, 130(7): 696–
Conference on Dry Mix Methods for Deep Soil Stabilization, 706. doi:10.1061/(ASCE)1090-0241(2004)130:7(696).
Stockholm, Sweden, 13–15 October 1999. A.A. Balkema, Cortellazzo, G., and Cola, S. 1999. Geotechnical characteristics of
Rotterdam, the Netherlands. pp. 125–153. two Italian peats stabilized with binders. In Dry Mix Methods for
Åhnberg, H., Bengtsson, P.E., and Holm, G. 2001. Effect of initial Deep Soil Stabilization, Proceedings of the International Con-
loading on the strength of stabilised peat. Ground Improvement, 5 ference on Dry Mix Methods for Deep Soil Stabilization,
(1): 35–40. doi:10.1680/grim.2001.5.1.35. Stockholm, Sweden, 13–15 October 1999. A.A. Balkema,
Al-Rawas, A.A. 2002. Microfabric and mineralogical studies on the Rotterdam, the Netherlands. pp. 93–100.
stabilization of an expansive soil using cement bypass dust and Earth Exploration, Inc. 1993. Preliminary geotechnical evaluation:
some types of slags. Canadian Geotechnical Journal, 39(5): 1150– Proposed improvements to Lindberg Road in West Lafayette,
1167. doi:10.1139/t02-046. Tippecanoe County, Indiana. West Lafayette City,. Indiana Internal
ASTM. 1989. Standard test method for one-dimensional consolida- report.
tion properties of saturated cohesive soils using controlled-strain Esrig, M.I., Mac Kenna, P.E., and Forte, E.P. 2003. Ground
loading. ASTM standard D4186. American Society for Testing stabilization in the United States by the Scandinavian lime cement
and Materials, West Conshohocken, Pa. dry mix process. In Grouting and Ground Treatment, Proceedings
ASTM. 2006. Standard practice for classification of soils for of the Third International Conference, New Orleans, La., 10–12
engineering purposes (Unified Soil Classification System). February 2003. Geotechnical Special Publication No. 120. Edited

Published by NRC Research Press


Bobet et al. 1115

by L.F. Johnsen, D.A. Bruce, and M.J. Byle. American Society of Johnsen, D.A. Bruce, and M.J. Byle. American Society of Civil
Civil Engineers, Reston, Va. pp. 501–514. Engineers, Reston, Va. pp. 575–585.
FHWA. 2000. An introduction to the deep mixing method as used in Lea, F.M. 2004. The chemistry of cement and concrete. 4th ed. Edited
geotechnical applications. Federal Highway Administration, Wa- by P. Hewlett. Butterworth-Heinemann, Oxford, UK.
shington, D.C. Document No. FHWA-RD-99-138. Liu, M.D., and Carter, J.P. 1999. Virgin compression of structured
Hayashi, H., and Nishimoto, S. 2005. Strength characteristics of soils. Géotechnique, 49(1): 43–57. doi:10.1680/geot.1999.49.1.43.
stabilized peat suing different types of binders. In Proceedings of Liu, M.D., and Carter, J.P. 2000. Modelling the destructuring of soils
the International Conference of Deep Mixing Best Practices and during virgin compression. Géotechnique, 50(4): 479–483. doi:10.
Recent Advances, Deep Mixing ’05, Stockholm, Sweden, 23–25 1680/geot.2000.50.4.479.
May 2005. Swedish Deep Stabilisation Research Centre, Linköp- Lorenzo, G.A., and Bergado, D.T. 2004. Fundamental parameters of
ing, Sweden. Vol. 1, pp. 55–62. cement-admixed clay – New approach. Journal of Geotechnical
Hebib, S., and Farrell, E.R. 2003. Some experiences on the and Geoenvironmental Engineering, 130(10): 1042–1050. doi:10.
stabilization of Irish peats. Canadian Geotechnical Journal, 40(1): 1061/(ASCE)1090-0241(2004)130:10(1042).
107–120. doi:10.1139/t02-091. Mesri, G., and Castro, A. 1987. Ca/Cc concept and Ko during
Hernandez-Martinez, F.G., and Al-Tabbaa, A. 2005. Strength secondary compression. Journal of the Geotechnical Engineering
properties of stabilised peats. In Proceedings of the International Division, ASCE, 113(3): 230–247. doi:10.1061/(ASCE)0733-
Conference of Deep Mixing Best Practices and Recent Advances, 9410(1987)113:3(230).
Deep Mixing ’05, Stockholm, Sweden, 23–25 May 2005. Swedish Pollard, S.J.T., Montgomery, D.M., Sollars, C.J., and Perry, R. 1991.
Deep Stabilisation Research Centre, Linköping, Sweden. Vol. 1, Organic compounds in the cement-based stabilization/solidifica-
pp. 69–78. tion of hazardous mixed wastes — Mechanistic and process
Hernandez-Martinez, F.G., and Al-Tabbaa, A. 2009. Effectiveness of considerations. Journal of Hazardous Materials, 28(3): 313–327.
different binders in the stabilisation of organic soils. In Proceed- doi:10.1016/0304-3894(91)87082-D.
ings of the International Symposium on Soil Mixing and Rao, S.N., and Rajasekaran, G. 1996. Reaction products formed in
Admixture Stabilisation, Okinawa, 19–21 May 2009. lime-stabilized marine clays. Journal of Geotechnical Engineering,
Holm, G. 2005. Keynote Lecture: Towards a sustainable society – 122(5): 329–336. doi:10.1061/(ASCE)0733-9410(1996)122:5
recent advances in deep mixing. In Proceedings of the Interna- (329).
tional Conference of Deep Mixing Best Practices and Recent Santagata, M.C., Bobet, A., Johnston, C., and Hwang, J.H. 2008.
Advances, Deep Mixing ’05, Stockholm, Sweden, 23–25 May One-dimensional compression behavior of a soil with high organic
2005. Swedish Deep Stabilisation Research Centre, Linköping, matter content. Journal of Geotechnical and Geoenvironmental
Sweden. Vol. 1, pp. K13–K24. Engineering, 134(1): 1–13. doi:10.1061/(ASCE)1090-0241(2008)
Hwang, J., Humphrey, A., Bobet, A., and Santagata, M. 2004. 134:1(1).
Stabilization and improvement of organic soils. Report No. Schnitzer, M., and Khan, S.U. 1972. Humic substances in the
FHWA/IN/JTRP-2004/38 for the Joint Transportation Research environment. Marcel Dekker, Inc., New York.
Program, Purdue University, West Lafayette, Ind. Sparks, D.L. 2003. Environmental soil chemistry, 2nd ed. Academic
Kang, Y.I. 2011. Stress–strain–strength behavior of a cement treated Press, New York.
clay. Ph.D. thesis (in progress), School of Civil Engineering, Townsend, D.C., and Klym, T.W. 1996. Durability of lime-stabilized
Purdue University, West Lafayette, Ind. soils. Highway Research Record, 139: 25–41.
Kang, Y.I., and Santagata, M.C. 2006. One-dimensional compression Tremblay, H., Leroueil, S., and Locat, J. 2001. Mechanical
behavior of cement-treated clay. In Ground Modification and improvement and vertical yield stress prediction of clayey soils
Seismic Mitigation, Proceedings of Sessions of GeoShanghai 6–8 from eastern Canada treated with lime or cement. Canadian
June 2006, Shanghai, China. Edited by A. Porbaha, S.-L. Shen, J. Geotechnical Journal, 38(3): 567–579. doi:10.1139/cgj-38-3-567.
Wartman, and J.-C. Chai. Geotechnical Special Publication USDA. 1998. Soil survey of Tippecanoe County. Natural Resources
No. 152. American Society of Civil Engineers, Reston, Va. pp Conservation Service, U.S. Department of Agriculture, Washing-
73–80. ton, D.C.
Kassim, K.A., and Clarke, B.G. 1999. Constant rate of strain Wissa, A.E.Z., Christian, J.T., Davis, E.W., and Heiberg, S. 1971.
consolidation equipment and procedure for stabilized soils. Consolidation at constant rate of strain. Journal of the Soil
Geotechnical Testing Journal, 22(1): 13–21. doi:10.1520/ Mechanics and Foundations Division, ASCE, 97(10): 1393–1413.
GTJ11312J. Yousuf, M., Mollah, A., Palta, P., Hess, T.R., Vempati, R.K., and
Lambrechts, J.R., Ganse, M.A., and Layhee, C.A. 2003. Soil mixing Cocke, D.L. 1995. Chemical and physical effects of sodium
to stabilize organic clay for I-95 Widening, Alexandria, VA. In lignosulfonate superplasticizer on the hydration of Portland
Grouting and Ground Treatment, Proceedings of the Third cement and solidification/stabilization consequences. Cement
International Conference, New Orleans, La., 10–12 February and Concrete Research, 25(3): 671–682. doi:10.1016/0008-8846
2003. Geotechnical Special Publication No. 120 Edited by L.F. (95)00055-H.

Published by NRC Research Press


Copyright of Canadian Geotechnical Journal is the property of Canadian Science Publishing and its content may
not be copied or emailed to multiple sites or posted to a listserv without the copyright holder's express written
permission. However, users may print, download, or email articles for individual use.

View publication stats

You might also like