Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Engineering Geology 51 (1998) 37–53

Effects of partial substitution of lime with ground granulated blast


furnace slag (GGBS) on the strength properties of lime-stabilised
sulphate-bearing clay soils
S. Wild a,*, J.M. Kinuthia a, G.I. Jones a, D.D. Higgins b
a Centre for Research in the Built Environment, University of Glamorgan, Pontypridd, Mid. Glamorgan CF37 1DL, UK
b Cementitious Slag Makers Association, Surrey, UK
Received 3 September 1997; accepted 26 June 1998

Abstract

Increasing global awareness of environmental pollution as well as increasing waste material disposal legislation is
providing impetus for material upgrading by stabilisation of in situ soil as an alternative to its export to land-fill and
replacement by imported granular fill. The use of ground granulated blast furnace slag (GGBS), an industrial
by-product, is well established as a binder in many cement applications where it provides enhanced durability,
including high resistance to chloride penetration, resistance to sulphate attack and protection against alkali silica
reaction (ASR). This paper reports on efforts to extend the use of GGBS to highway and other foundation layers by
determining the beneficial effect on strength of progressively substituting GGBS for lime in lime-stabilised clay soils,
particularly in the presence of gypsum. This paper describes the results of laboratory tests on lime-stabilised kaolinite
containing different levels of added gypsum and on lime-stabilised gypsum (selenite) bearing Kimmeridge Clay to
which, in both cases, the lime has progressively been substituted with GGBS. The tests determine the strength
development of compacted cylinders, moist cured in a humid environment at 30°C. The results illustrate that
substitution of lime with GGBS in stabilising gypsum-containing clays produces significant improvements in strength
development. The process has important practical applications, and the paper reports on-going pilot field trials aimed
at realising these applications. © 1998 Elsevier Science B.V. All rights reserved.

Keywords: Clay soils; Compressive strength; Highways; Lime; Slag

1. Introduction mechanical properties of the bulk material. These


properties will initially depend on the nature of
From the extensive research work reported on the immediate cation exchange reaction which
the pozzolanic activity between clay and lime, it is occurs, while the much slower pozzolanic reactions
apparent that the type and form of the reaction will determine the ultimate strength. Pozzolanic
products determine the physical, chemical and reactions occur over long periods (i.e. years) and,
depending on the curing environment, eventually
* Corresponding author. Fax: +44 1443 482169; cease altogether. Thus, the strength enhancing
e-mail: swild@glam.ac.uk reactions will also cease and deterioration may

0013-7952/98/$ – see front matter © 1998 Elsevier Science B.V. All rights reserved.
PII: S0 0 1 3 -7 9 5 2 ( 9 8 ) 0 0 03 9 - 8
38 S. Wild et al. / Engineering Geology 51 (1998) 37–53

even occur, depending on the chemical contributing silicates and aluminates is enhanced.
environment. The lime in the clay–lime mix will provide the
In clay–lime systems in general, the cementing required alkaline environment for slag activation
agent is calcium silicate hydrate (C–S–H ) gel and hydration, whilst also producing modification
(Croft, 1964; Sloane, 1965; Abdi, 1992; Wild et al., of the clay.
1993). The pore solutions of these systems contain In the presence of sulphates in the soil, expansive
silicate and aluminate ion species which are pro- reactions are likely to occur. Researchers have
duced by the dissolution of clay particles in the observed that the presence of SO2− ions in either
4
highly alkaline environment provided by the dis- PC, GGBS or PC–GGBS blends results in the
solved lime (i.e. OH− ions in solution). Since formation of a potentially expansive system mainly
kaolinite is high in alumina, the gel in kaolin- through the formation and subsequent growth of
ite–lime systems also contains alumina. In addi- ettringite (Mehta, 1973; Regourd, 1980). Similar
tion, crystalline calcium–aluminate hydrate expansive behaviour has been reported in case
(C–A–H ) phases (such as C AH and C AH ) studies involving clay–lime–sulphate systems
3 6 4 13
and calcium–aluminate silicate hydrate (Mitchell, 1986; Hunter, 1988; Snedker and
(C–A–S–H ) phases (such as C ASH ) are formed, Temporal, 1990).
2 8
as observed by Croft (1964). In the absence of Wild et al. (1993, 1996) have worked on clay–
excess alumina, only C–S–H phases would nor- lime systems with and without the presence of slag
mally form, as in the case of Portland Cement and/or gypsum. In lime-treated kaolinite clay in
(PC ) and silica fume ( Wild et al., 1995). the presence of gypsum, they observed nucleation
Microstructural evidence of similar gels in PC and and growth of ettringite crystals on the surface of
PC–GGBS blends suggests that the gel forms on clay particles. They also observed that clay–lime–
the surface of the solid particles (Richardson slag–gypsum systems can be expansive [as the
et al., 1994). principal reactants — Ca(OH ) , SiO , Al O ,
2 2 2 3
Introduction of slag into the clay–lime hydration SO2− and H O — are the same as those in PC
4 2
reaction will undoubtedly modify the clay–lime and PC–slag blends), but more importantly the
reaction products. Currently, an indication of the presence of slag in clay–lime–gypsum systems
manner in which these products may be modified lowers the deleterious expansions normally associ-
by slag can only be obtained from previous findings ated with ettringite formation. The present paper
on activated slag systems, as virtually no previous reports on the manner in which strength develop-
work has been carried out on clay–lime–slag sys- ment is modified when the lime in stabilised kaolin-
tems. The slag clearly provides additional alumina, ite–lime mixes is partially replaced by GGBS, in a
calcia, silica and magnesia to the system, depend- system which contains gypsum levels of 0, 2, 4 and
ing on the type and amount of slag (Regourd, 6% by weight. It also reports on the effects of
1980; Smolczyk, 1980). Since the principal reac- partially replacing lime with GGBS on the strength
tants introduced by slag are also present in the development of lime-stabilised Kimmeridge Clay,
clay–lime system as well as in PC–GGBS blends, the latter being a naturally occurring sulphate-
the strength of hydrated clay–lime–GGBS mix- bearing clay soil of Jurassic age.
tures is likely to be governed by the same factors
observed in hydrated PC–GGBS blends. These
factors include the properties of the C–S–H gel, 2. Materials
such as its amount, porosity and permeability;
fineness of all reactants involved; efficiency of 2.1. Kaolinite
mixing, temperature and curing time. Also, since
the slag introduces extra and more freely available Industrial kaolinite was used as a source of
alumina and silica compared to the clay–lime kaolinite. It was supplied by ECC International
system alone (due to its high reactivity in the Ltd, St. Austell, Cornwall, under the commercial
presence of lime), the formation of the strength- trade name ‘‘Standard Porcelain’’, as a white
S. Wild et al. / Engineering Geology 51 (1998) 37–53 39

Table 1 Table 3
Particle size distribution of ‘‘Standard Porcelain’’ ( ECC Engineering properties of ‘‘Standard Porcelain’’
International Ltd, 1987)
Consistency limits
Sieve size %
Liquid limit (%) 61
>53 mm 0.03 Plastic limit (%) 32
>10 mm 4 Plasticity index 29
<2 mm 70
Others

Table 2 Specific gravity 2.57


Chemical analysis and mineralogical composition of ‘‘Standard Maximum dry density (MDD) (Mg/m3) 1.50
Porcelain’’ ( ECC International Ltd, 1987) (Standard BS compaction)
Optimum moisture content (%) 26
Chemical analysis % pH 5.9
Natural moisture content (%) 2.5
SiO 48
2
Al O 37
2 3
Fe O 0.65 Table 4
2 3
TiO 0.02 Sulphate content of Kimmeridge Clay
2
CaO 0.07
MgO 0.30 Total sulphur 0.71%
KO 1.60 Sulphide (S2) 0.02%
2
Na O 0.10 Water soluble sulphate (SO ) 1.41%
2 3
Loss on ignition (LOI ) 12.5 Total sulphate (SO ) 1.73%a
3
Mineralogical composition a The total sulphate of 1.73% SO translates to 3.72% by weight
3
of gypsum (CaSO · 2H O).
4 2
Kaolinite 84
Micaceous material 13
Feldspar 1
other as a yellow grey (or brown) silty clay. The
Other minerals 2 material was sampled 1–3 m below ground level.
Borehole logs from this site report Kimmeridge
Clay as occurring at depths between 0.5 and 9 m
powder consisting of 84% kaolinite, 13% mica, 1% below ground level. The two types of clay men-
feldspar and 2% other minerals. Table 1 shows the tioned above have been identified as the same
particle size distribution of ‘‘Standard Porcelain’’, material, the change in colour being thought to
while Table 2 shows the chemical and mineralogi- result from localised weathering and oxidation
cal composition of the ‘‘as-supplied’’ material from (Littleton and Willavise, 1992). The clay also
ECC International Ltd. Table 3 shows the engi- contains pyrite [iron sulphide (FeS )] and fine
2
neering properties of Standard Porcelain. crystals of selenite (gypsum). The results of sul-
phate analysis, engineering properties and mineral
2.2. Kimmeridge Clay identification of the Kimmeridge Clay, which had
first been dried and crushed and fully homog-
Kimmeridge Clay was chosen for this study for enised, were carried out by Frodingham Cement
its proximity and similarity to Oxford Clay, which Company Ltd ( UK ) and are shown in Tables 4–6
was used in the construction of the M40 respectively. The presence of dickite, a type of
Oxford–Birmingham Motorway. It was obtained kaolinite, is significant as it makes the comparison
with the assistance and permission of Oxford City of test results with those of the model industrial
Council from a site near Oxford (an area of kaolinite (which also contains mica as a minor
farmland under development south east of phase) relevant. Thus, a study using kaolinite for
Oxford ). It occurred in two intermixed forms, one example might be expected to reflect the likely
as a firm to stiff dark (or blue) grey clay and the reactions in the Kimmeridge Clay and vice versa.
40 S. Wild et al. / Engineering Geology 51 (1998) 37–53

Table 5 Table 7
Engineering properties of Kimmeridge Clay Physical properties of lime (Buxton Lime Industries Ltd )

Consistency limits Trade (or product) ‘‘Limbux’’ hydrated lime


name
Liquid limit (%) 65 Chemical name Hydrated lime or calcium hydroxide
Plastic limit (%) 33 Physical form Dry white powder
Plasticity index 32 Melting/decomposition 580°C
temperature
Others Bulk density 480 kg/m3
Specific gravity 2.3 (water=1 at 4°C )
Maximum dry density (MDD) (Mg/m3) 1.57 Specific surface 300–1500 m2/kg
(Standard BS compaction) Solubility in water 1.76 g/l sat. sol. at 10°C
Optimum moisture content (%) 22 Solubility in Sugar solution
pH 9.6 other solvents
pH 12.4 (aqueous solution approx. 2 g/l )
Vapour pressure 0 at 20°C

Table 6 Table 8
Mineral phases in Kimmeridge Clay Chemical composition of lime (Buxton Lime Industries Ltd )
Mineral Chemical Compound Chemical Composition
formula formula (%)
Quartz SiO
2 Main (and hazardous) ingredient
Calcite CaCO
3 Hydrated lime Ca(OH ) 96.79
Ankerite Ca(Fe,Mg) (CO ) 2
32 (calcium hydroxide)
Dolomite CaMg(CO )
32
Gypsum (selenite CaSO · 2H O
4 2 Others
crystals) Calcite (calcium carbonate) CaCO 1.36
Muscovite (mica) ( K,Na)(Al,Mg,Fe) (Si Al )O (OH ) 3
2 3.1 0.9 10 2 Anhydrite (calcium sulphate) CaSO
4
0.06
Illite (mica) ( K,H O)Al Si AlO (OH )
3 2 3 10 2 Magnesia MgO 0.83
Dickite (kaolin) Al Si O (OH )
2 2 5 4 Ferric oxide Fe O
2 3
0.06
Anatase +Probable trace amount TiO
2 Alumina Al O
2 3
0.10
Silica SiO 0.46
2
Excess moisture HO 0.34
2

2.3. Lime Table 9


Particle size distribution of lime (Buxton Lime Industries Ltd )
Hydrated lime under the trade name ‘‘Limbux’’
was supplied by Buxton Lime Industries Ltd, Size (mm) % Passing
Buxton, Derbyshire; Tables 7 and 8 show its physi- 500 100.00
cal properties and chemical composition respec- 355 99.99
tively, while Table 9 shows its grading. 250 99.98
180 99.93
125 99.74
2.4. Slag 90 99.59
63 94.71
Ground granulated blast furnace slag (GGBS )
was supplied by Civil and Marine Slag Cement
Ltd, Llanwern, Newport; Table 10 shows its chemi- 3. Mix compositions and specimen preparation
cal composition and physical properties together
with a typical composition for Portland Cement For the purpose of sample preparation, it was
for comparison. found necessary to establish target dry density and
S. Wild et al. / Engineering Geology 51 (1998) 37–53 41

Table 10 sions 50 mm diameter and 100 mm length, were


Chemical composition and physical properties of GGBS and thoroughly mixed in a variable speed Kenwood
Portland Cement (Civil and Marine Slag Cement Ltd )
Chef mixer for 2 min before slowly adding the
Oxide Composition (%) calculated amount of water. Intermittent hand
mixing with palette knives was necessary to achieve
GGBS Portland Cement a homogeneous mix. The material was compressed
CaO 41.99 63 immediately after mixing into cylinders to the
SiO 35.34 20 prescribed dry density and moisture content. A
2
Al O
2 3
11.59 6 steel mould fitted with a collar, so as to accommo-
MgO 8.04 4.21 date all the material required for one sample, was
Fe O 0.35 3
2 3 used. Compaction, which was carried out immedi-
MnO 0.45 0.03–1.11
S 1.18 — ately after mixing, was achieved using a hydraulic
2
SO 0.23 2.3 jack. The prefabricated mould, which was oiled to
3
Insoluble residue 0.3 0.5 reduce friction, ensured that no further compaction
Relative density 2.9 3.15 was subjected to the soil once the desired volume
Bulk density (kg/m3) 1200 1400
Colour Off-white Grey
had been attained. The compacted cylinder was
Glass content #90 — left in the mould under pressure for a short period
before release, in order to allow for relaxation.
The densities would therefore be similar to those
moisture content values. Proctor compaction tests obtained using the standard Proctor test since the
were conducted in order to establish values of the dry density and moisture content remain constant.
maximum dry density (MDD) and optimum mois- The cylinders were extruded using a steel plunger,
ture (OMC ) for the range of compositions in each trimmed, cleaned of releasing oil, weighed, mea-
system and the mean MDD and OMC values were sured and wrapped in several runs of cling film.
adopted as the target values for the preparation The cylinders were labelled and placed in polythene
of test specimens. Samples were therefore expected, bags before being placed on a platform in sealed
within experimental error, to be of the same den- plastic containers. Water was always maintained
sity, volume and moisture content for all the below the platform to ensure that there was no
material compositions in a given system. In the evaporation from the samples. The plastic contain-
kaolinite–lime–GGBS–gypsum system, a dry den- ers were then placed in an environmental chamber
sity value of 1.41 Mg/m3 and a moisture content capable of maintaining temperatures to ±1°C and
value of 30% were adopted for all mixes (measured humidity to ±2% relative humidity. The samples
MDDs and OMCs varied between 1.40 and were moist cured for 7 and 28 days at 30°C and
1.46 Mg/m3 and 29–31% respectively). Similarly 100% relative humidity. Table 11 shows the details
for the Kimmeridge Clay–lime–GGBS system, a of the mix compositions used.
dry density value of 1.41 Mg/m3 and a moisture
content value of 24% was adopted (measured
MDDs and OMCs varied between 1.36 and 4. Testing and analysis
1.43 Mg/m3 and 21 and 28% respectively).
Although this means that, within each series, speci- At the end of the prescribed moist curing
mens of different composition will deviate slightly periods, samples were removed from the environ-
from their MDD and OMC values, the minor mental chamber. Any condensation on the cling
variation in void space which this will produce film covering the cylinders was removed with paper
would not be expected to have any significant tissue, prior to their being weighed. Two cylinders
effect on physical properties. per mix proportion were subjected to unconfined
Using the adopted mean density and moisture compressive strength ( UCS ) tests and the mean
content values, dry materials, enough to produce strength determined. A third test was sometimes
three compacted cylindrical test samples of dimen- necessary, especially when the strength variation
42 S. Wild et al. / Engineering Geology 51 (1998) 37–53

Table 11
Details of mix compositions

Lime (L) Slag (S) Gypsum (G) Moisture Dry density Mixes selected for
content (wt%) content (wt%) contenta (wt% SO ) content (%) (Mg/m3) use in XRD analysis
3
Kaolinite–lime– 6 0 0 30 1.41 ⻬
slag–gypsum
5 1 0 30 1.41 ⻬
3 3 0 30 1.41 ⻬
2 4 0 30 1.41 ⻬
1 5 0 30 1.41 ⻬
0 6 0 30 1.41 ⻬
6 0 1 30 1.41 ×
5 1 1 30 1.41 ×
3 3 1 30 1.41 ×
2 4 1 30 1.41 ×
1 5 1 30 1.41 ×
0 6 1 30 1.41 ×
6 0 2 30 1.41 ×
5 1 2 30 1.41 ×
3 3 2 30 1.41 ×
2 4 2 30 1.41 ×
1 5 2 30 1.41 ×
0 6 2 30 1.41 ×
6 0 3 30 1.41 ⻬
5 1 3 30 1.41 ⻬
3 3 3 30 1.41 ⻬
2 4 3 30 1.41 ⻬
1 5 3 30 1.41 ⻬
0 6 3 30 1.41 ⻬

Kimmeridge Clay– 5 0 1.73 24 1.41 ×


lime–slaga
3 2 1.73 24 1.41 ×
2 3 1.73 24 1.41 ×
1 4 1.73 24 1.41 ×
0 5 1.73 24 1.41 ×
6 0 1.73 24 1.41 ×
4 2 1.73 24 1.41 ×
2 4 1.73 24 1.41 ×
1 5 1.73 24 1.41 ×
0 6 1.73 24 1.41 ×
8 0 1.73 24 1.41 ×
5 3 1.73 24 1.41 ×
3 5 1.73 24 1.41 ×
2 6 1.73 24 1.41 ×
0 8 1.73 24 1.41 ×
10 0 1.73 24 1.41 ×
6 4 1.73 24 1.41 ×
4 6 1.73 24 1.41 ×
2 8 1.73 24 1.41 ×
0 10 1.73 24 1.41 ×

a It should be noted that the gypsum in the kaolinite–lime–slag system was added, while in Kimmeridge Clay it is inherent in the
clay as selenite crystals.
S. Wild et al. / Engineering Geology 51 (1998) 37–53 43

in the first two tests differed by more than 10% of observed subsequently to contain the greatest
the mean strength (normally a variation of ±10% amount of ettringite. In fact, specimens with low
of the mean was observed ). Before testing, the end lime–high slag contents tended to shrink during
surfaces of samples to be tested were gently moist curing rather than expand. Swelling of these
abraded to ensure a flat surface and good contact specimens during both moist curing and saturation
with the testing rig platens. A special self-levelling is addressed in a further paper ( Wild et al., in
device was used to ensure uniaxial load applica- press).
tion. An M30K JJ Instruments testing machine
capable of loading up to 30 kN was used to apply
the load at a compression rate of 1 mm/min. After 5. Results
testing, a small quantity of material was taken
from the interior of the tested sample for moisture 5.1. Kaolinite system
content determination to establish the moisture
condition at the end of each curing period. This Fig. 1(a–c) illustrates the effects (on the 7 and
was found to be within ±5% of the initial moisture 28 day UCS ) of progressive substitution of lime
content of the samples, and therefore any loss or with GGBS in kaolinite–6% by weight lime (con-
gain of moisture by the samples during curing was trol ) samples with and without added gypsum.
negligible (i.e. a moisture content of 30% would The levels of gypsum used (2, 4 and 6% by weight,
normally vary within 30±1.5%). based upon the clay and stabiliser combined) are
X-ray diffraction ( XRD) analysis was per- equivalent to approximately 1, 2 and 3% by mass
formed on selected mixes from the kaolinite– of SO .
3
lime–GGBS–gypsum system. The fewer mineral When no gypsum is present, for the 7 day curing
phases in the industrial kaolinite used compared period [Fig. 1(a–c)] partial substitution of lime
to those in the Kimmeridge Clay (see Section 2) with slag has very little effect on strength, produc-
made identification of new products simpler and ing only a slight increase in strength with increasing
was thus more effective in contributing to a clearer substitution. However, after 28 days the level of
understanding of the mechanisms involved in the substitution has a significant effect on strength,
interaction between the clay, lime, slag and producing an accelerating increase in strength with
gypsum. The selected mixes are indicated in increasing substitution of lime with slag until zero
Table 11. All X-ray diffraction runs were per- lime is reached, when strength falls sharply. This
formed using fixed settings for the X-ray equip- indicates that by 28 days the hydration of the slag
ment and a standard amount of powder specimen. is at a much more advanced stage with regard to
The work was carried out at the Department its contribution to strength than is the pozzolanic
of Mechanical, Materials and Manufacturing reaction between the lime and the kaolinite.
Engineering, University of Newcastle, Newcastle When gypsum is present, for the 7 day curing
upon Tyne, using a Philips diffractometer PW1965 period there is increasing strength enhancement,
and generator PW1730, a graphite monochroma- relative to the sulphate free material, with increas-
tor and Cu Ka radiation of wavelength l= ing substitution of lime with slag. This suggests
1.54179 Å. that the presence of gypsum not only accelerates
In addition to strength determination and X-ray the cementitious reactions with lime and kaolinite,
analysis, linear expansion measurements were also but also even more effectively accelerates the
carried out on these specimens. The dimensional hydration of the slag. After 28 days curing, there
changes exhibited by the specimens during 7 days is further strength enhancement by the gypsum,
of moist curing, which are the curing conditions but in this case the strength enhancement is much
prior to strength testing, were found to be relatively greater at low substitution levels and the enhance-
small (+2.1% to −0.55%). Also, specimens which ment becomes less as the lime is substituted with
exhibited the greatest swelling were those of high slag. Thus, at a 5:1 slag to lime ratio there is
lime–low slag contents, which were also those virtually no difference in the 28 day strengths of
44 S. Wild et al. / Engineering Geology 51 (1998) 37–53

ment when gypsum is present, particularly at the


higher SO levels, which suggests that the effect of
3
gypsum on the cementitious reaction between lime
and kaolinite is more pronounced at this later
stage. It should also be noted that, if only slag is
present and no lime (6:0), then strengths are
extremely low, there is no strength development,
and gypsum has no influence whatsoever on
strength (point B on graphs in Fig. 1). This sug-
gests that gypsum will only accelerate slag hydra-
tion if the slag is activated with lime, and will not
activate the slag on its own.
Fig. 2 presents the same 7 and 28 day strength
data, but in this case plotted against gypsum
content (as per cent SO ) rather than slag ( lime)
3
content. This again emphasises the increasing
strength enhancing effect of the sulphate additions
with increasing substitution of lime with slag, again
excepting 0% lime. In addition, it shows that there
is an optimum sulphate content of 1% SO for
3
maximum 7 day strength, which is principally a
result of acceleration of the slag hydration reac-
tions. For the composition 6L0S, sulphate content
has little beneficial effect on strength at 7 days.
After 28 days, however, sulphate content has a
very marked effect on the strength of this composi-
tion, which increases considerably in going from
0% SO to 2% SO , emphasising as previously
3 3
suggested that the reaction product resulting from
the clay–lime–gypsum reaction contributes sig-
nificantly to strength at these later periods.
Conversely, specimens containing a small amount
of lime and a large amount of slag (1L5S ) show
negligible variation in strength with increase in
sulphate level (at 28 days) simply because the
initial accelerating effect of the sulphate on slag
hydration is complete and the hydration level of
specimens without sulphate has reached that of
Fig. 1. Unconfined compressive strength vs. slag ( lime) content
specimens with sulphate.
for kaolinite–lime–GGBS cylinders with and without: (a) 1; (b) The way in which different levels of sulphate
2; (c) 3% SO equivalent of gypsum and moist cured for 7 and influence the rate of strength development is clearly
3
28 days at 30°C and 100% relative humidity. illustrated in Fig. 3(a–d). When no sulphate is
present [Fig. 3(a)], strength development of high
specimens with gypsum and without gypsum, irre- lime–low slag material (e.g. 6L0S) is very limited,
spective of the SO level. This indicates that the emphasising the very slow pozzolanic clay–lime
3
accelerating effect of the gypsum on the slag hydra- reaction. In contrast, strength development of the
tion is complete. However, at 0:6 slag to lime ratio low lime–high slag material (e.g. 1L5S) is substan-
there is, at 28 days, substantial strength enhance- tial, thus confirming the much more rapid activated
S. Wild et al. / Engineering Geology 51 (1998) 37–53 45

Fig. 2. Unconfined compressive strength vs. gypsum content (%SO ) for kaolinite–lime–GGBS cylinders with additions of varying
3
amounts of gypsum, moist cured for 7 and 28 days at 30°C and 100% relative humidity.

slag hydration reaction. When sulphate is intro- Fig. 3(c)]. Here the strength increase between 7
duced [see Fig. 3(b)], the 7 day strengths of low and 28 days for the high lime–low slag composi-
lime–high slag compositions increase substantially, tions is substantial, particularly when compared
whereas those of high lime–low slag compositions to that where no sulphate is present [Fig. 3(a)].
show much smaller increases, thus emphasising the This strength increase is attributed to formation
much greater accelerating effect of sulphate on of a cementitious reaction product which is not
activated slag hydration compared with its influ- formed in the absence of sulphate. It is suggested
ence on the clay–lime reaction. The effect of sul- that this product is ettringite.
phate on the activity of kaolinite–lime only mixes X-ray analysis was used principally for identifi-
is much more pronounced after 28 days [see cation of ettringite, lime and gypsum phases for
46 S. Wild et al. / Engineering Geology 51 (1998) 37–53

Fig. 3. Unconfined compressive strength vs. curing time for kaolinite–lime–GGBS cylinders with additions of varying amounts of
gypsum, moist cured for 7 and 28 days at 30°C and 100% relative humidity.

specimens cured at 1 and 4 weeks. The intensities ent [Fig. 6(a,b)], which indicates that either little
of these peaks relative to the kaolinite peaks give reaction had occurred or that the products of
an indication of the amounts present, although reaction were amorphous. The latter is undoubt-
this could be distorted by preferred orientation. edly the case, because cementation does indeed
After 1 and 4 weeks of moist curing of kaolinite– occur and the C–A–S–H gel formed due to reaction
6% by weight lime mixes, only X-ray diffraction of lime with kaolinite is not crystalline (Croft,
peaks due to mica, kaolinite and lime were appar- 1964; Sloane, 1965; Wild et al., 1993, 1996). The
S. Wild et al. / Engineering Geology 51 (1998) 37–53 47

crystalline products of the reaction which have various magnitudes of total stabiliser (TS ) content
been reported develop, if at all, at a very slow rate (i.e. lime+GGBS=5, 6, 8 and 10%). In contrast
(Croft, 1964; McCallister and Petry, 1992). When to the kaolinite–lime–gypsum system, where the
gypsum was present, the lime and gypsum peaks SO level was artificially varied, in Kimmeridge
3
present after 1 week of moist curing were not Clay the sulphate content is fixed at its as-sampled
present after 4 weeks of moist curing [Fig. 6(c,d )], level of 1.73% SO . Thus, for purposes of compari-
3
and new peaks identified as those of ettringite were son with the kaolinite–lime–gypsum system,
also visible after 4 weeks of moist curing. These Fig. 1(b) is the nearest equivalent (2% SO ). Also
3
ettringite peaks were present for mixes with little in the case of the Kimmeridge Clay, the total
or no substitution of lime with GGBS (i.e. mixes stabiliser content was varied from 5 to 10% in
of composition 6L0S06G and 5L1S6G). At higher order to establish, from a practical engineering
lime substitution (i.e. mixes of composition aspect, what an appropriate level of stabiliser
2L4S6G), no ettringite peaks were identifiable and would be. There are, with regard to strength
peaks of residual gypsum appeared [Fig. 6(f )]. development, clear similarities in behaviour
between the two clay systems [swelling of these
5.2. Kimmeridge Clay system specimens is addressed in a further paper ( Wild
et al., in press)]. Comparison of Fig. 4(b) with
Figs. 4 and 5 illustrate the effects (on the 7 and Fig. 1(b) shows parallel behaviour in the relation-
28 day UCS ) of progressive substitution of lime ship between 7 day and 28 day strength, in that at
with GGBS in Kimmeridge Clay–lime cylinders at low slag to lime ratios (6L0S) the 28 day strength

Fig. 4. Unconfined compressive strength vs. slag ( lime) content for Kimmeridge Clay–lime–GGBS cylinders containing a total
stabiliser ( TS) content ( lime+GGBS ) of: (a) 5; (b) 6; (c) 8; (d) 10% by weight, and moist cured for 7 and 28 days at 30°C and
100% relative humidity.
48 S. Wild et al. / Engineering Geology 51 (1998) 37–53

At both 7 days and 28 days of moist curing the


initial gradients of the strength versus binder com-
position curves [Fig. 4(a–d)] increase systemati-
cally with increase in total binder content. Also
the peak in strength versus composition curves
tends to move to higher slag/lime ratios as the
binder content increases. This suggests that at low
binder contents and high slag/lime ratios there is
insufficient total lime available in the system to
fully activate the slag. Fig. 5(a,b) suggests an
optimum slag/lime ratio of about 1.5 is required
to give maximum strength for the higher stabiliser
contents.
Another factor observed with Kimmeridge Clay,
not present with kaolinite, is that compositions
containing slag but no lime do show small strength
gains with time (points B and B∞ on the curves in
Fig. 4). This suggests that the Kimmeridge Clay
contains some component(s) which can produce
partial and limited activation of the slag. However,
this was not observed to occur in the kaolinite–
lime–gypsum system, which suggests that the com-
ponent producing this effect is something other
than gypsum alone.

6. Discussion

Enhanced strength development in clay–lime


and clay–lime–slag mixtures in the presence of
gypsum has been confirmed by many researchers
( Hasaba et al., 1982; Gupta and Seehra, 1989;
Fig. 5. Unconfined compressive strength vs. slag ( lime) content
for Kimmeridge Clay–lime–GGBS cylinders moist cured for 7
Abdi, 1992; Wild et al., 1993, 1996). In the current
and 28 days at 30°C and 100% relative humidity [ TS=total work, it was noted that the rate of strength devel-
stabiliser ( lime+GGBS) content]. opment was generally higher when gypsum was
present. Inclusion of gypsum in the clay–lime
is much greater than the 7 day strength and at system accelerates the strength development by a
high slag to lime ratios (1L5S) the 7 day and 28 second mechanism in addition to C–S–H gel for-
day strengths move closer to each other. Also, in mation, that is ettringite formation. This mode of
both cases the 7 day strength initially increases as strength enhancement results from the growth and
the slag to lime ratio increases, reaches a maximum development of crystalline ettringite [Fig. 6(d)].
and then decreases, although the maximum occurs The enhanced strength development, particularly
earlier with Kimmeridge Clay. The 28 day strength during the 7 to 28 day moist curing period, con-
developed at high slag–low lime compositions is firms the well-established accelerating effect that
attributed principally to the hydration of the slag gypsum has on the clay–lime reactions (Abdi,
which is activated by lime and initially accelerated 1992; Wild et al., 1993, 1996), particularly at later
by the presence of sulphate. stages. The acceleration is known to be due to
S. Wild et al. / Engineering Geology 51 (1998) 37–53 49

Fig. 6. X-ray diffraction traces of: (a,b) moist cured kaolinite–lime specimens without gypsum at 1 week and 4 weeks of moist curing
respectively; (c–f ) moist cured kaolinite–lime–GGBS specimens with 6% by weight of gypsum, at either 1 week and/or 4 weeks of
moist curing. (Note: Only peaks due to individual compounds are marked.)

rapid removal of aluminium from solution, by the formation of the cementitious calcium silicate,
formation of ettringite. aluminate and alumino-silicate hydrates.
When slag is present in clay–lime mixes, a major In the presence of gypsum, the sulphate further
factor determining the degree of formation of the promotes activation of the slag through lime–gyp-
cementitious hydration products is the hydration sum activation (Daimon, 1980; Smolczyk, 1980;
properties of the slag. The activation of slag by Wu et al., 1990; Gollop and Taylor, 1996). It
lime (and also in some cases by components in the accelerates and enhances slag hydration such that
clay soil ) promotes slag hydration and hence also strength is enhanced significantly by 7 days. Added
50 S. Wild et al. / Engineering Geology 51 (1998) 37–53

gypsum activates slag only in the presence of lime must result from hydration of the slag and the
(and/or any other alkali). Inclusion of gypsum in formation of more C–S–H gel, and not from
the clay–lime–GGBS system at high lime–low slag ettringite formation. Therefore, the increase in
compositions also accelerates the strength develop- strength as lime is substituted with slag is predomi-
ment due to the development and growth of nantly due to the increased formation of the
crystalline ettringite. This enhanced strength devel- cementitious silicate, aluminate and alumino-sili-
opment, particularly at low slag to lime ratios and cates (mainly of calcium). This will then lead to
high SO contents ( Fig. 3), is particularly apparent reduced porosity of the hydrated system (Bijen,
3
after 28 days of moist curing, which suggests that 1996) as the gel develops within the pore spaces
gypsum’s accelerating effect on the lime–kaolinite and closes off the capillary pores. Strength will
reaction is slower than on the slag hydration increase with increasing slag level as long as suffi-
reaction. cient lime is present to enable alkali–slag activa-
From Fig. 2(a,b), the optimal SO concentration tion. Therefore at low binder contents and high
3
for maximum strength development appears to slag/lime ratios, where insufficient lime is available
depend on the curing period. Obviously at pro- to fully activate the slag, strength decreases with
longed curing more sulphate will be consumed, increasing slag level. For example for Kimmeridge
resulting in an increase in optimum SO concen- Clay, when the total stabiliser ( lime+slag) does
3
tration and strength. However, there will clearly not provide enough lime for both cation exchange
be a limit to the amount of sulphate which can be and for pozzolanic activity to take place in addition
consumed, which will depend on the original lime to activation of the slag, any strength increase of
content. This limit, for the lime content employed clay–lime–slag mixes with increase in slag content
in the current work (6% by weight for the kaolinite appears to be limited to short curing periods of 7
system), and for 28 days of moist curing, would days [i.e. TS=5% in Figs. 4(a) and 5 (a)]. During
appear to be approximately 2% SO . this period, it is suggested that some proportion
3
XRD analysis results of clay–lime–GGBS– of the calcium ions are used for the cation exchange
gypsum mixes show that the favourable conditions while at the same time the short-lived alkalinity
for crystalline ettringite formation prevailing in only manages to activate the slag to a low level.
the kaolinite–lime–gypsum system are reduced Over longer periods (28 days), those samples with
significantly by the introduction of slag. In unsubstituted lime (i.e. low slag/lime ratios) will
PC–GGBS systems, it has been reported that the develop the highest strengths, the others being
higher amounts of C–S–H gel that are produced gradually denied hydration as the proportion of
readily absorb Al O (Gollop and Taylor, 1996), lime decreases, as shown in Fig. 5(b).
2 3
and possibly CaO, in a manner that these compo- Unlike in the case of kaolinite, there is significant
nents are not readily available for further reaction. strength development at total lime replacement
In the kaolinite–lime–slag–gypsum system at low with slag [points B and B∞ in Fig. 4(a–d)]. The 28
lime–high slag ratios where larger amounts of day strength at total lime replacement generally
C–S–H gel will be produced, it might be expected increases with increasing stabiliser content,
that less ettringite will be formed as more CaO whereas at 7 days there is little change in strength.
and Al O are bound in the C–S–H gel. Thus in There is no doubt therefore that in this case the
2 3
this case the pozzolanic reactions produce pre- slag is slowly activated. The activation must be
dominantly non-crystalline products and the due to one or more of the various mineral phases
strength enhancement is principally via C–S–H gel in Kimmeridge Clay. Two possible contenders
formation, the ettringite decreasing with increasing would be dolomite and calcite, both of which are
slag/lime ratio. Thus, the most significant effect sparingly soluble (dolomite less so than calcite)
that replacement of lime with slag has on the final and generate an alkaline environment in solution
product is the reduction or complete elimination (the pH of Kimmeridge Clay in solution was
of crystalline ettringite. Consequently, the strength observed as 9.6 compared to 5.9 for kaolinite).
enhancement in clay–lime–slag–gypsum mixes Thus, there is the possibility that some of the
S. Wild et al. / Engineering Geology 51 (1998) 37–53 51

phases shown in Table 6 help in activating the slag. range, the clay–lime reaction is more sensitive to
More research is definitely required here to estab- temperature than the slag hydration reaction,
lish the mechanism or the factors involved in this which further supports the case for using lime-
form of slag activation. activated slag stabilisation rather than lime
stabilisation.
Based on these findings, the Cementitious Slag
7. Practical applications Makers Association (CSMA), in collaboration
with the Transport Research Laboratory ( TRL)
Results from the current work indicate that high and other relevant industries such as the Buxton
(7 day and 28 day) strengths not previously achiev- Lime Industries (BLI ) ( UK ), have carried out a
able either with clay–lime or with clay–lime–gyp- pioneering highway pavement pilot field trial in
sum compositions can been achieved by partially the United Kingdom using GGBS on the A421
substituting the lime with ground granulated blast Tingewick bypass, west of Buckingham. The trial
furnace slag (GGBS ) in the stabilisation process. uses in situ soil stabilisation of sulphate-bearing
Further, results reported by Wild et al. (1996, in boulder clay using quicklime–Portland Cement–
pr) and Higgins et al. (in press) indicate that the GGBS optional compositions. The clay–lime–
compaction properties of both lime-stabilised GGBS option resulted in higher soaked and
kaolinite and lime-stabilised Kimmeridge Clay are unsoaked strength laboratory test results than
not significantly affected by the substitution of clay–lime–PC at equivalent stabiliser contents.
lime with GGBS, both in the absence and in the
presence of gypsum, and that both the material
plasticity and expansion potential are significantly 8. Conclusions and recommendations
reduced. These observations result in very signifi-
cant and potentially viable commercial and indu-
The following principal points may be deduced
strial applications, particularly for use in lime
from the results, with regard to the influence of
stabilisation of sulphate-bearing clay soils. At pre-
GGBS and gypsum on kaolinite and Kimmeridge
sent, there are no specific guidelines on the use of
Clay when stabilised with lime.
sulphate-bearing clay soils, and the most common
(1) Partial substitution of lime with GGBS gives
practice has been to avoid stabilisation of these
improved 7 day and 28 day strengths for both
soils, thus incurring heavy economic penalties in
kaolinite and Kimmeridge Clay. The maxi-
cutting, transporting and spoiling. Environmental
mum level of lime substitution is different for
factors, such as increasing global awareness of
pollution, and increasing legislation through taxes the two clay types. In the case of kaolinite,
on waste disposal methods, such as dumping, the effects are more pronounced in the pres-
encourage material upgrading by stabilisation of ence of gypsum.
in situ soil as an alternative to its export to land- (2) The most significant strength enhancement of
fill and replacement by imported granular fill. kaolinite stabilised with lime/GGBS over the
Thus, the lime stabilisation of sulphate-bearing first 28 days was either for (i) high lime–low
soils will result in global material resource savings slag mixes with gypsum, due to the contribu-
and better control of environmental degradation tion of gypsum to the longer-term kaolinite–
by waste material dumping. lime–gypsum reaction, or (ii) low lime–high
Curing at 30°C is an accelerated test, as in slag without gypsum due to the lime-activated
practice stabilisation occurs at ambient temper- slag hydration.
atures (normally 10–25°C ). Work carried out by (3) The greatest short-term strength enhancement
the authors ( Wild et al., 1998) on the shear was for low lime–high slag mixes with gypsum,
strength of slag/lime-stabilised Kimmeridge Clay due to the accelerating effect of gypsum on
at different temperatures (10–30°C ) has shown the lime-activated slag hydration.
that increase in temperature accelerates the rate of (4) Slag alone has no effect on both 7 day and 28
strength gain. In particular, in this temperature day strength of kaolinite while it does provide
52 S. Wild et al. / Engineering Geology 51 (1998) 37–53

limited but significant strength enhancement Bijen, J.G., 1996. Blast Furnace Slag Cement, Association of
in the case of Kimmeridge Clay. the Netherlands Cement Industry ( VNC ).
Croft, J.B., 1964. The pozzolanic reactivities of some New
(5) Substitution of lime by slag in stabilised South Wales flyashes and their application to soil stabiliza-
Kimmeridge Clay provides a maximum in 28 tion. In: Proc. ARRB, Australia, 1964, Vol. 2, Part 2 (Paper
day strength at a specific replacement level 120), pp. 1144–1167.
which depends on the total stabiliser content Daimon, M., 1980. Mechanism and kinetics of slag cement
(i.e. 2S4L at TS 6% by weight, 5S3L at TS hydration. In: 7th Int. Congress on the Chemistry of Cement,
8% and 6S4L at TS 10%). In the case of Paris, France, pp. III-2/1–III-2/9.
Gollop, R.S., Taylor, H.F.W., 1996. Microstructural and micro-
stabilised kaolinite, where only one total stabi- analytical studies of sulfate attack. V: Comparison of
liser content was investigated, the 28 day different slag blends. Cement Concrete Res. 26, 1029–1044.
strength continues to increase gradually up to Gupta, S., Seehra, S.S., 1989. Studies on lime-granulated blastf-
a replacement level of at least 5S1L, although urnace slag as an alternative binder to cement. Highway Res.
at 6S0L the strength falls sharply. Bull. 38, 81–97.
It is clear from the observations in this study Hasaba, S., Kawamura, M., Torii, K., 1982. Reaction products
and strength characteristics in the stabilised soil using desul-
that the amount of lime added, relative to the slag, furization by-product and blastfurnace slag. Trans. Jpn. Soc.
must be sufficient to activate the slag. Also, the Civ. Eng. 14, 251–253.
total stabiliser content (slag+lime) must be at the Higgins, D., Kinuthia, J.M., Wild, S., in press. Soil stabilisation
necessary level to provide the required bearing using lime-activated ground granulated blast-furnace slag [to
capacity and strength. Further work indicates that be presented and published in Proc. Sixth CANMET/ACI
the lime content must also, however, not be so Int. Conf. on Fly Ash, Silica Fume, Slag and Natural Pozzo-
lans in Concrete, Bangkok, Thailand, 31 May–5 June 1998].
great that excess lime is freely available over an Hunter, D., 1988. Lime-induced heave in sulphate bearing clay
extended period because the lime could react with soils. ASCE J. Geotech. Eng. 114, 150–167.
sulphates and cause swelling. Any expansive reac- Littleton, I., Willavise, G., 1992. Some observations on the use
tions involving sulphates must be exhausted during of lime to stabilise Kimmeridge Clay. Buxton Lime Indu-
the initial curing period prior to any possible stries Report, Ref. No. 91.
inundation by water. Work on the expansive McCallister, L.D., Petry, T.M., 1992. Leach tests on lime-
treated clays. Geotech. Test. J. 15 (2), 106–114.
behaviour of this material has been reported in a Mehta, P.K., 1973. Effect of lime on hydration of pastes con-
further paper ( Wild et al., in press). taining gypsum and calcium aluminates or calcium sulpho-
aluminate. J. Am. Ceram. Soc. 56, 315–319.
Mitchell, J.K., 1986. Practical problems from surprising soil
Acknowledgment behaviour. The twentieth Karl Terzaghi Lecture. ASCE
J. Geotech. Eng. 112, 274–279.
Regourd, M., 1980. Structure and behaviour of slag Portland
The authors would like to thank Civil and cement hydrates. In: 7th Int. Congress on the Chemistry of
Marine Slag Cement Ltd for supply of the slag. Cement, Paris, France, pp. III-2/10–III-2/26.
The authors would also like to thank Buxton Lime Richardson, I.G., Brough, A.R., Groves, G.W., Dobson, C.M.,
Industries and ECC International for supplying 1994. The characterisation of hardened alkali-activated
respectively the lime and the kaolinite. In addition, blast-furnace slag pastes and the nature of the calcium sili-
cate hydrate (C–S–H ). Cement Concrete Res. 24, 813–829.
the authors would like to thank Professor Richard Sloane, R.L., 1965. Early reactions in the kaolinite–hydrated
Neale, Head of School of the Built Environment lime–water system. In: Proc. 6th Int. Conf. on Soil Mechan-
and the technical staff in the School for provision ics and Foundation Engineering, Montreal, pp. 121–125.
of facilities and assistance. Smolczyk, H.G., 1980. Slag structure and identification of slags.
In: Proc. 7th Int. Conf. on the Chemistry of Cement, Paris,
France, pp. III-1/3–III-1/17.
Snedker, E.A., Temporal, J., 1990. M40 Motorway Banbury IV
References Contract — Lime Stabilisation. Highway and Transporta-
tion, December.
Abdi, M.R., 1992. Effects of calcium sulphate on lime stabilised Wild, S., Abdi, M.R., Leng Ward, G., 1993. Sulphate expansion
kaolinite. Ph.D. Thesis, University of Glamorgan, of lime-stabilised kaolinite. Part II. Reaction products and
Pontypridd. expansion. Clay Miner. 28, 569–583.
S. Wild et al. / Engineering Geology 51 (1998) 37–53 53

Wild, S., Sabir, B.B., Khatib, J.M., 1995. Factors influencing Education Funding Council for Wales (HEFCW ) Technol-
development of concrete containing silica fume. Cement ogy Foresight Initiative Project by the Construction Materi-
Concrete Res. 25, 1567–1580. als Research Unit, University of Glamorgan, April.
Wild, S., Kinuthia, J.M., Robinson, R.B., Humphreys, I., 1996. Wild, S., Kinuthia, J.M., Jones, G.I., Higgins, D.D., in press.
Effects of ground granulated blast-furnace slag (GGBS ) on Suppression of swelling associated with ettringite formation
the strength and swelling properties of lime stabilised kaolin- in lime-stabilised sulphate bearing clay soils by partial substi-
ite in the presence of sulphates. Clay Miner. 31, 423–433. tution of lime with ground granulated blast-furnace slag
Wild, S., Tasong, W.A., Veith, G., 1998. Utilisation of waste (GGBS). J. Eng. Geol.
materials in the stabilisation of land. Progress Report No. 2 Wu, X., Jiang, W., Roy, D.M., 1990. Early activation of slag
for the Progress Meeting with Industrial Partners, Higher cement. Cement Concrete Res. 20, 961–974.

You might also like