Sie sind auf Seite 1von 15

EGENEEPdNG

GEOLOGY
ELSEVIER Engineering Geology42 (1996) 223-237

Lime stabilization of clay minerals and soils


F . G . Bell
Department of Geology and Applied Geology, University of NataL Private Bag XIO, Dalbridge, 4014, South Africa
Received 12 April 1995; accepted 19 January 1996

Abstract

Clay soil can be stabilized by the addition of small percentages, by weight, of lime, thereby enhancing many of the
engineering properties of the soil and producing an improved construction material. In order to illustrate such
improvements, three of the most frequently occurring minerals in clay deposits, namely, kaolinite, montmorillonite
and quartz were subjected to a series of tests. As lime stabilization is most often used in relation to road construction,
the tests were chosen with this in mind. Till and laminated clay were treated in similar fashion. With the addition of
lime, the plasticity of montmorillonite was reduced whilst that of kaolinite and quartz was increased somewhat.
However, the addition of lime to the till had little influence on its plasticity but a significant reduction occurred in
that of the laminated clay. All materials experienced an increase in their optimum moisture content and a decrease in
their maximum dry density, as well as enhanced California bearing ratio, on addition of lime. Some notable increases
in strength and Young's Modulus occurred in these materials when they were treated with lime. Length of time curing
and temperature at which curing took place had an important influence on the amount of strength developed.

I. Introduction road and runway construction. Today stabilization


of clay soil by incorporation of lime is a technique
The addition of lime to soils to improve their widely used throughout the world to improve its
use for construction purposes has a very long use in construction. It is used in road construction
history. For instance, McDowell (1959) mentioned to improve sub-bases and subgrades, for railroad
that stabilized earth roads were used in ancient and airport construction, for embankments, as soil
Mesopotamia and Egypt, and that the Greeks and exchange in unstable slopes, as backfill for bridge
R o m a n s used soil-lime mixtures. More recently abutments and retaining walls, as canal linings,
the first tests involving soil stabilization were car- for improvement of soil beneath foundation slabs
ried out in the United States in 1904 (Clare and and for lime piles (Anon, 1985, 1990).
Cruchley, 1957). Lime was first used as a stabiliz- When lime is added to a clay soil it has an
ing agent of soil in m o d e m construction practice immediate effect on the properties of the soil as
in 1924 on short stretches of highway strengthened cation exchange begins to take place between the
by the addition of hydrated lime (McCaustland, metallic ions associated with the surfaces of the
1925). With the expansion of roads to cater for clay particles and the calcium ions of the lime.
the growth of m o t o r traffic in the 1930s, the use Clay particles are surrounded by a diffuse hydrous
of stabilization of soils began to increase. It was double layer which is modified by the ion exchange
extensively used during the Second World War for of calcium. This alters the density of the electrical

0013-7952/96/$15.00 © 1996 Elsevier Science B.V. All rights reserved


PII S0013-7952 (96) 00028-2
224 F (,. Bell~Engineering Geology 42 (1996) 223 237

charge around the clay particles which leads to Al203 and H is H/O) and calcium silicate alumi-
them being attracted closer to each other to form hate hydrate phases (C2ASH8, where S is SiO2)
flocs, the process being termed flocculation. It is may develop, especially when kaolinitic clays are
this process which is primarily responsible for the treated with lime.
modification of the engineering properties of clay The strength developed obviously is influenced
soils when they are treated with lime (Sherwood, by the quantity of cementitious gel produced and
1993). All types of clay minerals react with lime. consequently on the amount of lime consumed. In
The nature of the exchangeable cation does not fact, the amount of lime added should be related
make much difference in kaolinitic clay soils but to the clay mineral content of the soil. According
it can have a significant effect in clay soils contain- to Ingles (1987) a good rule of thumb in practice
ing montmorillonite. In fact, expandable clays tend is to allow 1% by weight of lime for each 10% of
to react readily with lime, losing plasticity immedi- clay in the soil. Exact prescriptions can be made
ately (Bell and Coulthard, 1990). after tests at and slightly to each side of this value.
When lime is added to a clay soil, it must first Because it is exceptional for the clay content of a
satisfy the affinity of the soil for lime, that is, ions soil to exceed 80%, it is normally not necessary to
are adsorbed by clay minerals and are not available add more than 8% lime.
for pozzolanic reactions until this affinity is satis- Extended curing times and elevated temper-
fied. Because this lime is fixed in the soil and is atures promote pozzolanic reactions. For instance,
not available for other reactions, the process has significantly improved strength can be developed
been referred to as lime fixation (Hilt and with relatively small increases in temperature.
Davidson, 1960). The lime fixation point corres- Conversely, if the temperature falls below around
ponds with the point where further addition of 4~C, pozzolanic reactions are retarded and may
lime does not bring about further changes in the cease at lower temperatures. In fact, pozzolanic
plastic limit. This therefore is the optimum addi- reactions may remain dormant during periods of
tion of lime needed for maximum modification of low temperatures to regain reaction potential when
the soil and is normally between 1 and 3% lime temperatures increase.
added, by weight. Beyond this point lime is avail-
able to increase the strength of the soil.
In addition to cation exchange, reaction occurs
between the silica and some alumina of the lattices 2. Clay minerals and the addition of lime
of the clay minerals, especially at the edges of clay
particles. In other words, the highly alkaline envi- Clay soils have a wide range of mineralogical
ronment produced by the addition of lime gives composition. They may consist of various propor-
rise to the slow solution of alumino-silicates which tions of different types of clay minerals, notably
are then precipitated as hydrated cementitious kaolinite, illite, mixed-layer clays and montmoril-
reaction products (Diamond and Kinter, 1966). lonite, of non-clay minerals, notably quartz, and/or
These reaction products contribute to flocculation organic matter and colloidal matter. Very small
by bonding adjacent soil particles together and as amounts of certain clay minerals may exert a large
curing occurs they strengthen the soil. Such pozzo- influence on the physical properties of a clay
lanic reactions are time dependent, strength devel- deposit. In addition, the degree of crystallinity is
oping gradually over a long period. For maximum important, clay minerals with poorly-ordered crys-
reactivity, the pH value of the pore fluids in the tallinity have different properties from those with
voids should remain at around 12.4. The solubility well-ordered crystallinity. The different properties
of silicon and aluminium ions is very high at this of the various families of clay minerals can be
value. Calcium silicates form as long as the highly explained partly by the different levels of activity
alkaline conditions persist. In addition, small on the surface of the clay particle. Expansive clay
amounts of calcium aluminate hydrate phases minerals such as montmorillonite exhibit a high
(e.g., C4AH13 and C3AH6, where C is CaO, A is cation exchange capacity, whereas non-expansive
F.G. Bell~Engineering Geology 42 (1996)223-237 225

Table 1
Composition and average mechanical properties of kaolinite, montmorillonite and quartz

Kaolinite Montmorillonite Quartz

1. Chemical composition
SiO2 47.0% 60.4% 99.19%
AI20a 38.3% 19.2% 0.2%
Alkali (K20, Na20) 0.8% 1.1% -
Fe203 0.5% 1.7% 0.01%
TiO2 - - 0.04%
CaO - 1.9% 0.07%
MgO - 5.6% Trace
Loss of ignition 13.3% 8.4% 0.35%
pH at 10% solids 5.0_+0.5 8.6+0.5 5.8_+0.3
Cation exchange capacity (meq/100 g) 2 16 80-120
2. Physical properties
Particle size distribution
0.06-0.002 mm 6% 24% 32%
Less than 0.002 mm 94% 76% 68%
Specific gravity 2.57 2.63 2.65
Liquid limit (%) 75 114 52
Plastic limit (%) 42 67 38
Plasticity index (%) 33 47 14
Activity 0.35 0.47 0.27
Unconfined compressive strength (kPa) a 350 127 318
Modulus of elasticity (MPa) 35 26 24
Optimum moisture content (%) 29.5 20.0 28.5
Maximum dry density (Mg/m a) 1.40 1.29 1.41
CBR 1 9 1

aAt optimum moisture content.

clay m i n e r a l s like k a o l i n i t e h a v e a relatively low rillonite reduces its plasticity. Plasticity is n o t


c a t i o n exchange capacity. necessarily r e d u c e d w h e n lime is a d d e d to k a o l i n i t e
D e t e r m i n a t i o n o f the c a t i o n e x c h a n g e c a p a c i t y a n d a n y changes are n o t as significant as they are
o f k a o l i n i t e a n d m o n t m o r i l l o n i t e , b o t h with a n d with m o n t m o r i l l o n i t e . This different b e h a v i o u r o f
w i t h o u t different a d d i t i o n s o f lime, was c a r r i e d k a o l i n i t e with r e g a r d to plasticity p r e s u m a b l y is
o u t using the m e t h o d p r o p o s e d b y G a l l a w a y a n d e x p l a i n e d b y the difference in c a t i o n exchange.
B u c h a n a n (1964). T h e c h e m i c a l c o m p o s i t i o n s o f T h e c a t i o n exchange c a p a c i t y was c o m p a r e d
the k a o l i n i t e a n d m o n t m o r i U o n i t e which were used
in this testing p r o g r a m m e are given in Table 1.
Table 2
T h e results o f these tests are given in Table 2 f r o m Results of cation exchange determinations
which it can be seen t h a t the c a t i o n exchange
c a p a c i t y o f m o n t m o r i l l o n i t e was r e d u c e d w h e n Lime content (%) Cation exchange capacity (meq/100 g)
lime was a d d e d b u t at 6% a d d i t i o n , b y weight, it
b e g a n to increase. O n the o t h e r h a n d , the c a t i o n Kaolinite Montmorillonite
exchange c a p a c i t y o f k a o l i n i t e rose initially with 0 4 88
2% a d d i t i o n o f lime, then declined a n d r e m a i n e d 2 12 80
m o r e o r less s t e a d y with a d d i t i o n s o f 4% o r more. 4 9 56
These changes seem to occur a r o u n d the p o i n t o f 6 9 66
8 10 62
lime fixation. T h e initial r e d u c t i o n o f c a t i o n ion
10 10 73
exchange c a p a c i t y on a d d i t i o n o f lime to m o n t m o -
226 F.G. Bell~Engineering Geology 42 (1996) 223-237

with the development of strength as increasing


amounts of lime were added to kaolinite and
montmorillonite. There was little relationship
between the two parameters which suggests that
cation exchange capacity is not an important factor
in terms of strength development when lime is
added to these clay minerals.
Samples of kaolinite and montmorillonite were
mixed with 2, 4, 6, 8 and 10% lime at optimum
moisture content and compacted at maximum dry
density, according to the standard Proctor com-
paction test. They then were sealed in polythene Fig. 2. SEM photomicrograph of montmorillonite with 4%
containers and allowed to cure at 20°C for one added lime.
year. This extended curing time was to allow for
the growth of reaction products. The samples then silicon, then followed aluminium. Traces of
were prepared and examined with an electron calcium and, to a lesser extent, potassium were
microscope. When kaolinite was treated with lime, present. The calcium presumably was part of the
it would appear that the edges of the particles reaction products formed by attack of lime upon
were attacked by lime, this accounting for their the kaolinite. The elements present in the montmo-
ragged appearance ( Fig. 1). In addition, individual rillonite-lime mixtures were primarily silicon and
particles appeared to be bonded together with gel- aluminium, and to a lesser extent calcium and
like material which coated the flocs. This would magnesium.
agree with the work of Eades and Grim (1960), X-ray diffraction was used in an attempt to
and of Croft (1964). The addition of lime to identify the reaction products formed when lime
montmorillonite gave rise to small hazy patches of was added to the clay material. When kaolinite
reaction products at the edges of the particles was treated with varying increments of lime, the
(Fig. 2). intensities of the characteristic kaolinite peaks were
Constitutive elemental analysis was carried out both decreased and broadened. The decreased
by attaching an energy dispersive spectrometer to intensities would suggest an attack by lime on the
the scanning electron microscope. This focuses an kaolinite particles (Eades and Grim, 1960). The
electron beam on a point on the sample. X-ray broadening of the peaks suggests either a slight
photographs generated by the sample are con- expansion of some of the layers or a change in the
verted into electrical impulses and the chemical orientation of the particles, as would perhaps occur
elements present are determined. In the case of due to the flocculating effects of the lime on the
kaolinite plus lime, the major element present was kaolinite. New reflections were visible at d-spacings
of 3.87, 3.67, 3.035, 1.619 and 1.582 A. However,
these new peaks only appeared in samples having
a lime content exceeding that of the lime fixation
point. The first new reaction product appeared to
be calcite, CaCO3. This would probably result
from carbonation of the lime. Other new products
which appeared were calcium silicate hydrates in
the form CSHI (tobermorite) and C 3 8 2 H 3 (afwil-
lite) and calcium aluminate hydrates of the form
C4AH13 CAH10 and CaAH11. These results agreed
with those of previous workers such as Croft
(1964) and Diamond and Kinter (1965) who
Fig. 1. SEM photomicrograph of kaolinite with 4% added lime. reported the presence of CAH and CSH reaction
F. G. Bell~EngineeringGeology 42 (1996) 223-237 227

products. Eades and Grim (1960) also reported (CaOH), 93.85%; calcium carbonate (CaCO3) ,
the formation of CSH and maintained that the 5.00%; silicon dioxide (SiO2), 0.63%; magnesium
cause of this was lime eating into the kaolinite oxide (MgO), 0.34%; aluminium oxide (A1203),
particles around their edges. A series of new X-ray 0.09%; ferric oxide (Fe203), 0.05%; calcium sul-
peaks were found by Willoughby et al. (1968) but phate (CaSO4), 0.04%; loss of ignition, 3.14%
the only one which resembles those of the present
study was that at 1.58,~ which corresponds 3.1. Consistency limits
to either CAHlo (1.60) or C352Hlo (1.589).
Abdelkader (1981) noted the presence of In most cases the effect of lime on the plasticity
hydrogarnets of the form CaASnH6_zn, but no of clay soil is more or less instantaneous, the
evidence of these were found in the present study. calcium ions from the lime causing a reduction in
Similarly calcium aluminium hydrates of the form plasticity, the soils becoming more friable and
CAH, C4AH13 or CAH~o would appear to form more easily worked. The clay particles undergo
when montmorillonite is treated with lime, along flocculation to form aggregates. The aggregates
with calcium silicate hydrates of the form CSH. behave like particles of silt. Very small quantities
In a recent study of lime stabilization by Arabi of lime are required to bring about these changes
and Wild (1989) of Devonian red marl, the authors in plasticity. Generally the amount needed varies
disputed the formation of calcium silicate hydrate from 1 to 3% depending on the amount and type
(CSH) reaction products. They maintained that of clay minerals present in soil.
the gel produced was a calcium silicate aluminate The plastic limit of montmorillonite increased
hydrate. This marl contained illite, quartz and up to 4% addition of lime. Thereafter it declined
feldspar with minor amounts of chlorite and hema- slightly (Fig. 3). Above 2% addition of lime, the
tite. Because of the difference in the composition plastic limit of kaolinite suffered a slight reduction.
of the material tested, then this may account for Quartz underwent an increase in plastic limit up
the different reaction products produced. to 2% lime added, after which it remained more
Quartz also was treated with additions of lime or less steady or declined a small amount (Fig. 3).
and examined by X-ray diffractometry. A new
peak occurred in the range 3.704-3.766 .A for all llO

lime contents while another occurred in the range I O0

3.145-3.160 ,~ for lime contents of 4% and above. 9O

These reaction products were identified as calcium


silicate hydrates of the form C3S2H3. Eades et al. - -'~........~"~'"~'"'.g2..........~ ............*'
11; 7O
(1962) also reported the formation of CSH in iii

lime-quartz mixtures. ~. . - "

• ......
•-'" ...... .0.......11-............
-~ 40,
3. Lime stabilization, clay minerals and engineering ~3o
behaviour :~20

10
Three of the major constituents in clay soils are I I I I I
0
kaolinite, montmorillonite and fine-grained quartz. 0 2 4 6 8 10
LIME CONTENT (%)
Consequently a series of tests were carried out on
these minerals to assess the influence upon them LIQUID LIMITS PLASTIC LIMITS
of the addition of various amounts of lime. The o---=~ Kaolinite ~ Kaolinite
composition of these three minerals used in the C ~Montrnonllonite 0 - - - - 0 Montmonllonlte
0 ....... 0 Quarlz • ...... • Quartz
present study, as well as their average physical
properties are outlined in Table 1. The composition Fig. 3. Plastic and liquid limits of kaolinite, montmorillonite
of the lime was as follows: calcium hydroxide and fine quartz with various additions of lime.
228 F.G. Bell/Engineering Geology 42 (1996) 223 237

When kaolinite was treated with 2% lime it 3.2. Compaction


underwent an increase of some 28% in liquid limit.
Croft (1964) suggested that such a rise was due to The addition of lime to clay materials increases
the action of hydroxyl ions which modified the their optimum moisture content and reduces their
affinity of the surfaces of the clay particles for maximum dry density for the same compactive
water. Further additions of lime gave rise to little effort (Table 3). The reduction in dry density could
change in the liquid limit. Taken together with the be due to an immediate formation of cementitious
values of plastic limit, the resultant plasticity index products which reduce compactibility and hence
of kaolinite showed an overall increase. By con- the density of the treated soil (Lees et al., 1982).
trast, the value of liquid limit of montmorillonite Obviously the design of lime-stabilized soil must
decreased with increasing lime content (Fig. 3). have due regard for these changes in compaction
Hence the action of hydroxyl ions, referred to optima, which in fact are proportional to the
above, would appear not to be significant in expan- amount of lime added (more lime, lower maximum
sive clays. The reduction in liquid limit along with dry density and higher optimum moisture content).
the increase in plastic limit, produced a consider- However, the strength gain in the soil will normally
able reduction in the plasticity index of montmoril- more than compensate for the changes in compac-
lonite. The liquid limit of quartz rose sharply with tion optima, and they should not be regarded as
small additions of lime. However, the plasticity of disadvantageous.
quartz also increased with the addition of lime. Higher compaction densities are obtained in
The shrinkage characteristics of expansive clayey lime-treated kaolinitic clays, than in soils in which
soils are improved significantly by the addition of clay minerals with expandable lattices predominate
lime (Fig. 4). The addition of lime to such clay (Table3). Croft (1964) suggested that this was
soils also reduces, or indeed removes, their poten- due to the higher demand of expandable clay
tial for swelling. The reduction in swelling is minerals for adsorbed and lubrication water than
believed to be mainly due to substitution of other that of kaolinite. Of the three minerals tested,
cations by calcium. The reduction in swelling also quartz developed the highest compaction densities.
means that there is a decrease in moisture absorp- The California bearing ratio (CBR) of clay soils
tion in lime-treated soils (Bhasin et al., 1978). is improved by the addition of lime (Table 3).
Indeed the CBR increases immediately after the
18~ addition of lime and continues to increase with
time if there is lime available in excess of the lime
;s-- fixation point.

12--
3.3. Strength and deformation

< The amount of strength increase in a clay soil


that can be produced by adding lime is dependent
on the pozzolans present. When the desirable
6-- pozzolans are available, they react readily with
lime to improve the strength of soil-lime mixtures.
3-- Nonetheless it would appear that the absolute
amount of silica or alumina required to sustain
pozzolanic reaction in soils is relatively small.
I I 3 ~ -7
Hence clays generally show a significant increase
0 2 4 6 8 10
in strength when lime is used for stabilization.
LIME CONTENT (%)
Expansive clays respond more quickly to
Fig. 4. Influence of lime content on the linear shrinkage of strength increase. For instance, montmorillonite
montmorillonite. showed a rapid initial increase in unconfined corn-
F. G. Bell~Engineering Geology 42 (1996) 223-237 229

Table 3
Comparison of the results of compaction and California bearing ratio tests carried out on kaolinite, montmorillonte and quartz with
those obtained when these materials were treated with optimum lime contents

Material Optimum lime Optimum moisture Maximum dry C B R (%)


content (%) content (%) density ( M g / m 3)

Kaolinite 0 29 1.4 1
6 31 1.33 14
Montmorillonite 0 20 1.29 9
4 25 1.15 18
Quartz 0 28 1.41 1
6 32 1.40 22

pressive strength with small additions of lime content, and excessive addition of lime reduces
(Fig. 5). Its lime strength optimum was around strength (Fig. 5). This is due to the fact that lime
4% compared with that of kaolinite which varied itself has neither appreciable friction nor cohesion.
between 4 and 6%, whilst that of quartz ranged Among the different variables affecting the
between 4 and 8%. The addition of small amounts strength of lime-stabilized clay soil, curing is of
of lime to kaolinite and quartz also give rise to major importance. Its effect on strength is a func-
notable increases in unconfined compressive tion of time, temperature and relative humidity
strength (Figs. 6 and 7). Furthermore, montmoril- (Mitchell and Hooper, 1961). The strength
lonite mixed with a low lime content attained increases rapidly at first, generally during the first
maximum strength in less time than one to which 7 days of curing, then increases more slowly at a
a higher content of lime had been added. Indeed more or less constant rate. For example, with an
strength does not increase linearly with lime optimum lime content of 6% kaolinite developed

900
1100
~ 8OO I000

700 9oo
w Z
800
~ 6oo
700
.° >
N 5oo .

6OO
~ 4oo
0
~ 3oo 8 400 :"~1
LIJ
Z
@
,-r 200 3 0 0 ~F
7"
0
IO0
0 I I I I I
o I I I I I
0 2 4 6 8 10 2 4 6 B 10
LIME C O N T E N T (%) LIME CONTENT [%]

o ............ ~ 1 day strength O ............ ~ 1 aay strength


o.-----o 3 day strength o-----o 3 day strength
o- ..... o 7 day strength o- ..... o 7 day strength
o . . . . . . o 14 day strength o ...... o 14 clay strength
o o 28 d a y strength ¢ ~ 28 d a y strength

Fig. 5. Unconfined compressive strength of montmorillonite Fig. 6. Unconfined compressive strength of kaolinite with vari-
with various additions of lime. ous additions of lime.
230 F G. Be///Engineering Geology 42 ~1996) 223-237

"6 5000 declined after 28 days to less than that of montmo-


rillonite alone. Such decreases can be quite signifi-
cant and A1-Rawi (1981) noted that at times they
z 4000 may exceed 30%. They are associated with con-
0 ...... 0 ...... 0 siderable decreases in dry density. Hence it is
3000
important to determine the optimum lime content
,0 " " "" * for mix design. It also is important to achieve a
(1- homogeneous soil-lime mixture, otherwise large
./- . ' t i ~ i ! " ~ ' -- • "0 . . . . . "0 . . . . . "0
o 2000
variations in strength due to deviation of lime
0
n
content from optimum will occur as a result of
L.kl .." ~........0 ............ O............. '0,,., '0 non-uniform distribution of lime.
1000
z
The optimum lime content represents the point
© of lime saturation and may move, according to
z
0
I I I I I clay mineral content, towards higher increments
2 4 6 8 10 with increasing length of curing time. For instance,
LIME CONTENT (%) this optimum and its associated strength generally
tend to increase with longer curing for montmoril-
o .......... o 1 clay strength
lonitic soils. On the other hand, in the case of
o-----~ 3 d a y strength
kaolinitic soils the optimum lime content does not
o-. -. -o 7 d a y strength
vary significantly with curing period.
o ..... o 14 d a y strength Kaolinite showed a rapid increase in Young's
c o 21 d a y strength modulus with additions of between 2 and 4% lime
which brought about increases of several hundred
Fig. 7. Unconfined compressive strength of quartz with various
additions of lime.
percent. Above 4% addition there was little
increase in Young's modulus and indeed it declined
somewhat (Fig. 8). Similarly montmorillonite
an unconfined compressive strength of 1.05 MPa underwent an initial increase in Young's modulus
after 28 days of curing which represented an with up to 4% addition of lime, after which there
increase of 3.5 times compared with the untreated was a general decline. Additions of up to 6 or 8%
unconfined compressive strength of this mineral. of lime gave rise to steady increases in the value
This steady slow gain in strength in kaolinitic clay of Young's modulus of fine quartz (again of several
soils provides an appreciable factor of safety for hundred percent), after which the value remained
designs based on 28-day strength. With the addi- more or less constant.
tion of 4% lime to fine quartz, the material rose The value of Young's inodulus increased with
in strength from 300 kPa untreated to 5.0 MPa increasing period of curing for kaolinite and
after 28 days of curing, an increase of some quartz. However, the maximum value of Young's
16.7 times. modulus of montmorillonite appeared to be
As can be inferred from the above, the uncon- reached after 14 days, after which a gradual decline
fined compressive strength of lime-treated soil occurred. Holm (1979) also found that lime stabili-
develops rapidly with increasing addition of lime zation of clay soil increased the value of Young's
until an optimum lime content is reached beyond modulus with increasing length of curing period,
which the strength continues to increase at a tbr instance, by some 15 times after 3 weeks from
reduced rate or begins to decline. The latter is treatment and around 35 times after 16 months.
exhibited by montmorillonite. The optimum addi-
tion of lime to montmorillonite (4%) gave rise to
4. Lime stabilization of glacial deposits, Teesside,
an increase in strength of around 1.75 times that
northeast England
of the untreated material. However, with additions
of 6% and above, the unconfined compressive Much of the area around Teesside is covered
strengths of such soil-lime mixtures actually with deposits of late Pleistocene age, namely, tills
F. G. Bell~Engineering Geology 42 (1996) 223-237 231

110
100
,a
90
80
u 70
~"...................... O"'"
60
50
40

C~ 30 /
o
20
10

0
I I I I I
0 2 4 6 8 10
LIME CONTENT (%)

O..........o 1 day strength


o--------o 3 day strength
o-. - . -o 7 day strength
o ..... o 14 day strength ~.~ AHuviurn ~ Rock outcrop

21 day strength f~ Boulder Clay "~ - 82' shoreline

~-:.~ Laminated C~ay ,I,A Marginal sand


Fig. 8. Young's modulus of kaolinite with various additions
0 1 2 3 4 5 km
of lime.
Scale

and laminated clay (Fig. 9). The laminated clays Fig. 9. Distribution of Boulder Clay and Tees Laminated Clay
overlie the tills, the latter being separated into in the Teesside area.
upper and lower divisions by sands and gravels.
In other words, there is a tripartite division of the ral mineralogy of the fine fraction of the Upper
tills into Lower Boulder Clay, Middle Sands and Boulder Clay and Tees Laminated Clay are as
Upper Boulder Clay. These tills are regarded as of follows:
late Devensian age. The tills were products of
successive ice sheets. The first deposited lodgement 4.1.1. Upper Boulder Clay
till and outwash as it retreated and these deposits Illite, 25-40%; kaolinite, 20-30%; chlorite,
subsequently were overridden by a second ice sheet >5%; quartz, 5-35%; calcite, >5%; dolomite,
which left behind its own lodgement till and out- >5%.
wash. As this second ice sheet retreated, drainage
eastward from the Teesside area was blocked, 4.1.2. Tees Laminated Clay
which led to the formation of a proglacial lake. Illite, 17-43%; kaolinite, 23-34%; chlorite,
The Tees Laminated Clay was deposited in this 9-19%; quartz, 4-26%; calcite, 2-7%; dolomite,
proglacial lake and rests directly upon the Upper 3-7%.
Boulder Clay. The Upper Boulder Clay varies in consistency
from firm to hard, generally being stiff to very
4.1. A brief review of geotechnical properties stiff. Its natural moisture content may be as low
as 10% or as high as 20%, but generally is between
Block or core samples were obtained from the 12 and 14%, with a bulk density between 1.90 and
Upper Boulder Clay and the Tees Laminated Clay. 2.22 Mg/m 3, with a mean value of 2.12 Mg/m 3.
X-ray diffraction analysis suggested that the gene- This till is characteristically unsorted, consisting
232 F.G. Bell~Engineering Geology 42 (1996) 223- 237

of a clast and fine fraction. The clast fraction 4.2. Stabilization with lime
usually accounts for less than l0 or 20% of the
deposit. Hence it is a matrix dominated till. Clay Some of these deposits, notably the laminated
size material often constitutes over 50% of the fine clays, can give rise to problems in earthworks. One
fraction, silt around 20% and sand the remainder. of the primary reasons for this is their loss of
However, sandy tills do occur. The till is of low strength on remoulding, that is, they tend to be
to medium plasticity, with plasticity indices com- medium sensitive to sensitive. Hence a programme
monly within the range 18 33%. The plastic and of testing was carried out to determine the extent
liquid limits tend to vary between 13 and 18%, to which the geotechnical properties of these
and 29 and 42%, respectively. The results of quick deposits could be enhanced by the addition of lime.
undrained triaxial tests indicate a wide range of Samples of Upper Boulder Clay and Tees
cohesion, from 52 to 522 kPa, with values most Laminated Clay were mixed with varying amounts
frequently between 150 and 250 kPa. Drained tri- of lime to determine to what extent this enhanced
axial tests tend to give effective values of cohesion their engineering properties. As soil stabilization
and angle of friction around 65-80kPa and is used most frequently in relation to the formation
11-18°, respectively. of subgrades and subbases in road construction,
The natural moisture content of the Tees the soils were subjected to related tests, namely,
Laminated Clay tends to be higher than that of plastic and liquid limit tests, compaction tests,
the Upper Boulder Clay, ranging from 15 to 45%, California Bearing Ratio tests and strength tests.
but it generally is between 20 and 33%. As a Since satisfactory mixing is important in soil stabi-
lization, the lime was thoroughly mixed with the
consequence its bulk density is lower ranging from
soil samples by means of a rotary mixer.
1.51 to 2.03 Mg/m 3, with an average value of 1.87
Mg/m 3. As far as its particle size distribution is
4.2.1. Consistency
concerned, the Tees Laminated Clay can contain
As noted above, the addition of lime to a clay
up to 10% fine sand, with the silt content varying
soil normally has an instantaneous effect. In the
between 27 and 43%. Hence, clay is the dominant
case of the Upper Boulder Clay, however, the
particle size, comprising between 44 and 76%.
addition of lime had little influence on plasticity.
Although the plasticity of the Tees Laminated Clay
This was because lime tended to increase both the
varies from low to very high, reflecting the different
plastic limits and liquid limits by similar amounts
proportions of silt and clay, it generally possesses (Table 4). The largest change in plasticity took
a medium to high plasticity. Bell and Coulthard place with only 2% addition of lime.
(1991) found that values of plastic and liquid Turning to the Tees Laminated Clay, both its
limits were between 18 and 31%, and 29 and 78%, plastic and liquid limits, as well as plasticity, are
respectively, with respective average values of 26 higher than those of the Upper Boulder Clay. The
and 56%. Linear shrinkage usually amounts to addition of lime to this clay brought about a
around 9 to 12%. In terms of consistency this clay notable reduction in its plasticity, which was
ranges from soft to stiff, with most material belong- mainly the result of increases in plastic limit,
ing to the firm category. When the plasticity index although the liquid limit did decline in value with
is plotted against percentage clay fraction on an increasing addition of lime (Table 4).
activity chart, the Tees Laminated Clay would The addition of lime reduced the linear shrink-
appear to have a medium to high potential for age of both these soils (Table4). Increasing
expansion. The average shear strength of this clay, amounts of lime gave rise to increasing reductions
as obtained form undrained triaxial tests, is 60 kPa, in shrinkage, the most noticeable reductions being
although it does exhibit a wide variation from 20 attained with small additions of lime. With the
to 102 kPa. The clays are medium sensitive to addition of 8% lime the amount of reduction in
sensitive and tend to undergo a notable loss of linear shrinkage of both soils is more or less
strength when remoulded. similar.
F.G. Bell~EngineeringGeology42 (1996)223-237 233

Table 4
Values of plastic limit (PL), liquid limit (LL) plasticity index (PI) and linear shrinkage (LS) of Upper Boulder Clay and Tees
Laminated Clay treated with various amounts of lime

Soil Property Amount of lime added (%)

0 2 4 6 8

Upper Boulder Clay PL (%) 14 25 23 21 18


LL (%) 30 42 40 41 37
PI (%) 16 17 17 20 19
LS (%) 6 2 1 1 1
Tees Laminated Clay PL (%) 26 36 34 33 31
LL (%) 58 57 53 50 49
PI (%) 32 19 19 17 18
LS (%) 10 4 3 2 2

4.2.2. Compaction and California bearing ratio soils in wetter than original condition to be com-
Compaction and California bearing ratio tests pacted satisfactorily.
were carried out on untreated and treated samples
of both soils, the treated samples being compacted 4.2.3. Strength and elasticity
with optimum strength additions of lime. Both Clay soils usually exhibit significant increases in
soils underwent a slight reduction in maximum strength when treated with lime. As noted above,
dry density and increases in optimum moisture the strength of such soil mixtures is influenced by
content when treated with lime (Table 5). Both the character of the soil, the amount of lime
soils showed appreciable increases in California added, the length of time available for curing and
bearing ratio when lime was added to them the conditions under which this takes place (i.e.,
(Table 5). temperature and humidity), moisture content, and
Unlike cement, lime treatment tends to flatten the time elapsed between mixing and compaction.
the compaction curve, in this way ensuring that a The strength and modulus of elasticity of both
certain proportion of prescribed density can be the Upper Boulder Clay and the Tees Laminated
achieved over a much wider range of moisture Clay were increased with addition of lime. The
contents (Fig. 10). Hence more relaxed moisture maximum strength was achieved in the Upper
control specifications are possible with lime than Boulder Clay with the addition of 6% and in the
cement. In addition, the optimum moisture content Tees Laminated Clay with the addition of 4% lime.
tends to move towards higher values, enabling These additions also gave the highest values of

Table 5
Values of (a) compaction and California bearing ratio (CBR) tests on Upper Boulder Clay and Tees Laminated Clay treated with
optimum amounts of limea, and (b) unconfined compressivestrength (UCS) and Young's modulus (E) of the Upper Boulder Clay
and Tees Laminated Clay treated with various amounts of lime and cured at 20°C for 7 days

Soil Amount Compaction CBR Property Amount of lime added (%)


added (%)
Optimum moisture Maximumdry 0 2 4 6 8
content (%) density (Mg/m3)

Upper Boulder 0 18 1.81 9 UCS (kPa) 538 762 1056 1597 1452
Clay 6 20 1.75 24 E (MPa) 35 49 56 58 52
Tees Laminated 0 22 1.65 5 UCS (kPa) 178 584 889 776 847
Clay 4 25 1.60 19 E (MPa) 15 21 43 38 40

~Quantities at which highest unconfined compressive strength achieved after 7 days curing.
234 F.G. Bell~Engineering Geology 42 (1996) 223-237

(a) in strength was approximately four and a half


1.90
times and whereas in the Upper Boulder Clay it
1,89
was less than three times).
1.88
1.87 -
The length of time involved in curing and the
amount of water used in mixing also were varied,
1.86

1.85
-
/ , ....e"........... ...~. \,, the results being given in Table 6, from which it
- /I
/ ° "°SS %%% ~ can be seen that, with the exception of those
1.84 -
• .o" samples mixed with only 10% water, in almost all
Z 1,83 -
,,"°° "@',,i
~ the other cases there is a general rise in strength
1.82 -
with increasing length of curing time. Generally
1.81 - the most notable increases in strength occur within
/ the first 7 days when cementitious reactions are
1.80 --F .... I I I -- I r. . . . l
most active. The amount of water available for
14 15 16 17 18 19 20 21
hydration and reaction to form cementitious bonds
MOISTURECONTENT (%) influences the strength which can be attained
(Sabry and Parcher, 1979). A low mix water
1.70 -
(b) content means that not only is insufficient water
j,/f -'l,,., available for efficient compaction but also that
water is rapidly used in the hydration process and
1.65-
/ q so the maximum gain in strength is developed after
iJ ~%
a short period of time. In order to develop maxi-
/ %
/s %
1.60-
Table 6
or)
Values of unconfined compressive strengths (kPa) of Upper
Boulder Clay and Tees Laminated Clay treated with various
Z 1.,'~ - amounts of lime for different curing times and different mix
a .. ",,,, water contents

Lime content Moisture content Duration(days)


1,50- (%) (%)
0 7 14 28

A. Upper Boulder Clay


1,45
2 10 502 597 549 467
10 14 18 22 26 30 20 708 796 862 972
MOISTURECONTENT (%) 30 401 441 568 761
4 10 614 710 704 598
20 712 935 1252 1497
30 492 627 794 917
MOISTURECONTENT(%]
6 10 576 849 871 734
11----O 0%
20 821 1206 1429 1622
O..... .O 3%
30 594 876 951 1123
= • 6%
B. Tees Laminated Clay
Fig. 10. Examples of compaction curves of (a) Upper Boulder 2 10 349 362 345 238
Clay and (b) Tees Laminated Clay samples containing 0, 3 and 20 396 484 557 602
6% lime. 30 170 183 160 195
4 10 476 639 516 45l
20 621 723 801 876
Young's modulus. Nonetheless the proportionate 30 321 362 407 487
increases in strength were greater in the Tees 6 10 509 525 323 222
20 570 669 724 880
Laminated Clay than the Upper Boulder Clay (in
30 311 331 261 379
the Tees Laminated Clay the maximum increase
F. G. Bell/Engineering Geology 42 (1996) 223-237 235

mum strength it is necessary to use a mix water 3.0


content slightly in excess of optimum moisture
content. This allows achievement of a satisfactory 2.5
I
dry density with sufficient moisture available to ¢
I

ensure full hydration. If the mix water content is l


I

well in excess of the optimum moisture content, l


l
d
then a softer soil mixture is produced but this will
continue to increase in strength with increasing
length of curing period. /
I
Higher temperatures accelerate curing, resulting ,.o I
f
in higher strengths. Mateous (1964), for instance,
found that specimens of clay soil cured at 35°C
developed twice the strength or more of those
specimens cured at 25°C. Similar results have been
reported by Arabi and Wild (1989). The influence 0 1 2
of curing temperature on the development of (a) UME(%)
strength in these tills and laminated clays is 3.5 7
illustrated in Fig. 11. By doubling the curing tem-
perature from 20 to 40°C, the strength developed
3.0 .....O
by soil-lime mixtures is increased substantially. In
those samples with lower amounts of additive the 2.5 fj ''''It ~
strength more or less doubles and in most others
it increases by approximately 1 MPa. Fig. 12 indi-
cates that gains in strength are greater at higher 2.01 ]r'j''/'wf
temperatures, the rise being quite marked above
30°C.

zw
1.0 /
5. Conclusions
o
0.5
Many of the important engineering properties
of clay soils are enhanced by the addition of lime.
All the same, the properties of such soil-lime 0 . 0 ~
I ] I I I ! I l
mixtures vary and depend upon the character of 0 I 2 3 4 5 6 7 8
the clay soil, the type and length of curing, and LIME (%)

the method and quality of construction.


(b) o----o At 40°C c = At 20"C
When lime is added to clay soils, calcium ions
initially are combined with or adsorbed by clay Fig. 11. The influence of temperature on the unconfined com-
minerals. These changes continue up to the lime pressive strength of (a) Tees Laminated Clay and (b) Upper
fixation point. This addition of lime contributes Boulder Clay, mixed with various amounts of lime and cured
for 7 days.
towards the improvement of soil workability but
not to an increase in strength. Lime added in
excess of the fixation point is utilized in the cemen- be obtained by adding small quantities of lime.
tation process and the effectiveness of lime stabili- The largest increases are in montmorillonitic clays,
zation is dependent upon the development of whilst the plastic limits of kaolinitic clays change
reaction products formed from the attack of lime to a lesser extent. Also the plastic limits of fine
on the minerals in a deposit of clay. quartz does not change significantly. In the tests
Increases in the plastic limits of clayey soils can carried out both the liquid limits of kaolinite and
236 F.G. Bell/Engineering Geology 42 (1996) 223-237

4.0- lime. This means that the plasticity index of the


Tees Laminated Clay is significantly reduced by
3.5 lime. That of the Upper Boulder Clay is slightly
increased by the addition of lime. Linear shrinkage
of each soil is reduced by the addition of lime.
As far as compaction is concerned, the addition
of lime to both the till and the laminated clay
decreases their maximum dry density and increases
their optimum moisture content. The latter trend
Zw u jZ 2.0- is more than offset by the increases in strength
which are developed by soil-lime mixtures. The
C) 1.5. California bearing ratio is increased in both soils.
-~z i J //'/'~' The strength of these two soils is considerably
enhanced by the addition of lime. The optimum
gain in strength appears to be with 4-6% lime.
The gain in strength is influenced by the amount
0.5 of mix water, the length of time of curing and the
1.o 2'o 2o temperature at which curing takes place. The high-
TEMPERATURE o C est gains in strength occur when the mix water
e----e Upper Boulder 13--ID Tees Laminated content is just in excess of the optimum moisture
Cloy Cloy
content. Strength continues to increase with
Fig. 12. The influence of curing temperature on the development increasing length of curing time but generally the
of unconfined compressive strength. Samples cured for 7 days. most notable increases occur within the first 7
days. Temperature of curing has a significant
quartz increased with the addition of lime, while influence on the strength attained, and above 30oc
that of montmorillonite was reduced. The plasticity the increase in strength may be dramatic.
of montmorillonitic clay is reduced but that of
kaolinitic clay need not be. The susceptibility to
volume change of expansive clays also is reduced. References
Lime treatment increases the strength of clay
materials. Montmorillonitic clays respond much Abdelkader, M.O., 1981. Lime-Soil Stabilization, Unpublished
more rapidly to lime stabilization, and so exhibit Ph.D. Thesis, Birmingham University, England.
earlier gains in strength than do kaolinitic clays. A1-Rawi, N.M., 1981. The effect of curing temperature on lime
However, after further ageing, the additional stabilization. Proc. 2nd Aust. Conf. on Engineering Materi-
strength achieved by montmorillonitic clay is not als, Sydney, 611-662.
Anon, 1985. Lime Stabilization Construction Manual. Eighth
as high as that achieved by kaolinitic clays. The Edition. National Lime Association, Arlington, VA.
most significant increases in strength occur when Anon, 1990. Lime Stabilization Manual. British Aggregate
lime is added to fine quartz. Young's modulus of Construction Materials Industry, London.
clay materials is increased by the addition of lime. Arabi, M. and Wild, S., 1989. Property changes induced in clay
The Upper Boulder Clay of the Teesside area is soils when using lime stabilization. Mun. Engr., 6:85 99.
Bell, F.G. and Coulthard, J.M., 1990. Stabilization of clay soils
generally a stiff or very stiff clay of low plasticity with lime. Mun. Engr., 7:125 140.
while the Tees Laminated Clay is generally soft to Bell, F.G. and Coulthard, J.M., 1991. The Laminated Clays of
firm and of medium to high plasticity. Normally Teesside. In: M.G. Culshaw, J.C. Cripps, J.A. Little and C.F.
the till is stronger than the laminated clay. Moon (Editors), Quaternary Engineering Geology, Engi-
The plastic limits of both soils is increased by neering Geology Special Publication No 7, The Geological
Society, London, pp. 339 348.
the addition of lime, as is the liquid limit of the Bhasin, N.K., Dhawan, P.K. and Mehta, H.S., 1978. Lime
till. However, in the case of the Tees Laminated requirement in soil stabilization. High. Res. Board, Bulletin
Clay the liquid limit declines with the addition of No 7, Washington, DC, 15-26.
F G, Bell/Engineering Geology 42 (1996) 223-237 237

Clare, K.E. and Cruchley, A.E., 1957. Laboratory experiments (Editor), Ground Engineer's Reference Book. Butterworths,
in the stabilization of clays with hydrated lime. Geotech- London, 38/1-38/26.
nique, 7: 97-110. Lees, G., Abdelkader, M.O. and Hamdani, S.K., 1982. Effect
Croft, J.B., 1964. The processes involved in the lime stabiliza- of the clay fraction on some mechanical properties of lime-
tion of clay soils. Proc. Aust. Road Res. Board, 2, Part 2: soil mixtures. J. Inst. High. Engr., 11: 3-9.
1169-1203. Mateous, M., 1964. Soil-lime stabilization of soils for highway
Diamond, S. and Kinter, E.B., 1965. Mechanisms of soil-lime purposes - final report. Illinois High. Engng. Series, No. 25.
stabilization. High. Res. Rec., 92: 83-102. McCaustland, D.E.J., 1925. Lime dirt in roads. Proc. Natl.
Diamond, S. and Kinter, E.B., 1966. Adsorption of calcium Lime Assoc., 7: 12-18.
hydroxide by montmorillonite and kaolinite. J. Colloid Inter- McDoweU, C., 1959. Stabilization of soils with lime, lime-flyash
face Sci., 22: 240-249. and other lime reactive materials. High. Res. Board, Bull.
Eades, J.L. and Grim, R.E., 1960. Reaction of hydrated lime 231, Washington, DC, 60-66.
with pure clay minerals in soil stabilization. High. Res. Mitchell, J.K. and Hooper, D.R., 1961. Influence of time
Board, Bull. 262, Washington, DC, 51-63. between mixing and compaction on properties of lime stabi-
Eades, J.L., Nichols, F.P. and Grim, R.E., 1962. Formation of lized expansive clay. High. Res. Board, Bull. 304, Washing-
new minerals with lime stabilization as proven by field experi- ton, DC, 14-31.
ments in Virginia. High. Res. Board, Bull. 335: Washington, Sabry, M.M.A. and Parcher, J.V., 1979. Engineering properties
DC, 31-39. of soil-lime mixes. Proc. ASCE J. Trans. Eng. Div., 197,
Gallaway, B.M. and Buchanan, S.J., 1964. Lime stabilization TEl, 25-35.
of clay soils. Proc. Conf. Aust. Road Res. Board, 2: Sherwood, P.T., 1993. Soil Stabilization with Cement and Lime:
1169-1203. State-of-the-Art Review. Transport Research Laboratory,
Hilt, G.H. and Davidson, D.T., 1960. Lime fixation of clayey Her Majesty's Stationery Otfice, London.
soils. High. Res. Board, Bull. 262, Washington, DC, 20-32. Willoughby, D.R., Gross, K.A., Ingles, O.G., Silva, S.R. and
Holm, G., 1979. Lime column stabilization - experience con- Spiers, V.M., 1968. The identification of reaction products
cerning strength and deformation properties. Vag-och Vot- in alkali stabilized clays by electron microscopy, X-ray and
tenbyggaren, 25(7-8): 45~19. electron diffraction. Proc. Aust. Road Res. Board, 4:
Ingles, O.H., 1987. Soil stabilization, Chapter 38. In: F.G. Bell 1386-1408.

Das könnte Ihnen auch gefallen