You are on page 1of 17

Soils and Foundations 2014;54(4):760–776

HOSTED BY The Japanese Geotechnical Society

Soils and Foundations

www.sciencedirect.com
journal homepage: www.elsevier.com/locate/sandf

Settlement performance of pad footings on soft clay supported


by stone columns: A numerical study
Micheál M. Killeena,1, Bryan A. McCabeb,n
a
Arup Consulting Engineers, Ringsend Road, Dublin, Ireland
b
College of Engineering and Informatics, National University of Ireland, Galway, Ireland
Received 14 March 2013; received in revised form 10 April 2014; accepted 22 April 2014
Available online 28 July 2014

Abstract

The settlement behaviour of small loaded areas (such as pad and strip footings) on soft soil supported by stone columns is poorly understood.
The lack of confinement associated with peripheral columns and the sharp stress decay with depth are features which are incompatible with the
widely-used unit cell method for infinite column groups thereby rendering small group behaviour particularly difficult to analyse. Useful field
data is virtually non-existent and although innovative laboratory modelling has been insightful, the findings cannot easily be extrapolated to field
scale. In this study, a 3-D finite element analysis in conjunction with an elastic–plastic soil model is used to identify the effect of variables in the
design process and interactions between them: these include column arrangement, spacing, length, and Young's modulus of the column material.
A simplified method is proposed to relate the settlement of small groups to a reference unit cell settlement predicted by current analytical
approaches.
& 2014 The Japanese Geotechnical Society. Production and hosting by Elsevier B.V. All rights reserved.

Keywords: Footings; Stone columns; Settlement improvement factor

1. Introduction stiffness and strength of the compacted stone serve to


relieve the soil of some of its load. Stone columns can be
It is well established that vibro-replacement stone col- used in a variety of scenarios ranging from small loaded
umns provide an effective means of increasing the bearing areas (e.g. footings) to wide loaded areas (e.g. floor slabs,
capacity and reducing the compressibility of soft soils. The embankments).
benefits arise from the fact that there is partial replacement The optimum deployment of stone columns beneath
of the host soil by a more competent material; the enhanced small loaded areas (such as pad and strip footings) is
arguably the most challenging aspect of stone column
settlement prediction in soft soils. Sexton et al. (2013)
n
Corresponding author. Tel.: þ353 91 492021; fax: þ 353 91 494507. have used 2-D axisymmetric Finite Element (FE) modelling
E-mail address: bryan.mccabe@nuigalway.ie (B.A. McCabe).
1
Formerly of the College of Engineering and Informatics, National
to demonstrate that some recent analytical methods (Castro
University of Ireland, Galway, Ireland. and Sagaseta, 2009 and Pulko et al., 2011 in particular)
Peer review under responsibility of The Japanese Geotechnical Society. are successful in capturing the key factors influencing

http://dx.doi.org/10.1016/j.sandf.2014.06.011
0038-0806/& 2014 The Japanese Geotechnical Society. Production and hosting by Elsevier B.V. All rights reserved.
M.M. Killeen, B.A. McCabe / Soils and Foundations 54 (2014) 760–776 761

the performance of infinite column grids, which can be the strength and stiffness of the column material. A simple
implemented mathematically using a unit cell model. design method predicting the settlement of groups of
However, unit cell models cannot capture the behaviour floating (partial depth) or end-bearing (full depth) columns
of peripheral group columns beneath footings that are not is proposed, based upon geometrical corrections to unit cell
equally confined on all sides. The smaller the group, the settlement predictions.
more influential these peripheral columns become. Further-
more, the increment of vertical stress beneath footings 2. General finite element modelling details
decays much more acutely with depth than that beneath
widely loaded areas, allowing partial depth treatment to be 2.1. Column group variables
used in practice even though analytical methods are poorly
developed in this respect. PLAXIS 3D Foundation FE software (Version 2.2) has
A database of measured field settlement improvement been used in this study to capture the three-dimensional
factors in fine-grained soils compiled by McCabe et al. nature of small groups of columns supporting pad footings.
(2009) highlights a dearth of such data for strip and pad The settlement performance of various configurations of
footings. High quality physical models of footings on soft columns was investigated by varying the number of columns
clay supported by stone columns (i.e. Muir Wood et al., (N), column spacing (s) on a square grid (with one
2000; McKelvey et al., 2004; Black et al., 2011; Sivakumar exception) and column length (L). The combination of N
et al., 2011; Shahu and Reddy, 2011) have been informa- and s dictated the footing width (B). The diameter of stone
tive, although there are obvious difficulties in extrapolating columns (d) is normally not a design variable as the poker is
model test findings to field scale, and the proportion of the of fixed diameter (430 mm used for the bottom feed system)
area under each footing that has been replaced with stone in and a column diameter of d ¼ 600 mm was adopted for the
some of these tests has tended to lie at the high end of what subsequent Finite Element Analysis (FEA) which is a
might commonly be used in practice. Research in which the typical average diameter of columns formed in soft soils.
finite element method has been used to model ground The area replacement ratio or area ratio is a parameter which
improved with stone columns relates to wide area loading measures the proportion of in situ soil replaced with stone
in the main, using either unit cell (e.g. Domingues et al., columns. For a finite group of columns, the area ratio is
2007a) or 2-D axisymmetric (e.g. Elshazly et al., 2008a) defined as the ratio of the footing area (A) to the total area of
approximations. Some 3-D modelling of large column grids columns beneath the footing (AC):
has also been carried out (i.e. Gäb et al., 2008); however,
A 4B2
other than Kirsch (2008) no 3-D modelling of column- ¼ ð1Þ
supported footings, carried out in conjunction with an AC Nπd2
advanced constitutive soil model, has been published. In However, for an infinite grid of columns, the area ratio is
this paper, PLAXIS 3D Foundation is used to model the defined as the ratio of the area of the zone of influence of one
behaviour of rigid square pad footings supported by various column (A ¼ s2 for a square grid) to the area of a single column
stone column configurations. The soil profile at the former (AC):
UK geotechnical test site at Bothkennar, Scotland, is used
A 4s2
as the basis of the modelling as it is representative of many ¼ 2 ð2Þ
soft soil profiles internationally for which the applicability AC πd
of stone columns is of growing interest. This research The column length was increased in 1 m increments in
identifies the roles and interactions of key factors relevant the subsequent FEA from L ¼ 0 m (i.e. an untreated footing)
to the settlement design of small groups of stone columns, to L ¼ 13 m to examine the settlement performance corre-
such as column arrangement, spacing and length, as well as sponding to partial depth treatment. The final analysis at

Fig. 1. Typical layout of a 3  3 group of columns beneath a 3 m square footing in the Bothkennar soil profile (see Section 4.1).
762 M.M. Killeen, B.A. McCabe / Soils and Foundations 54 (2014) 760–776

L ¼ H ¼ 13.9 m (H ¼ depth of the Carse clay modelled) 2.2. Column configurations


represents a full length stone column founded on a rigid
layer. The various column group parameters are illustrated 2.2.1. Influence of column confinement
for a 3  3 group of columns beneath a 3 m square footing The degree of lateral support or confinement experienced
in Fig. 1. by columns in a group is dependent upon the number of
The influence of other design parameters such as column columns and the area ratio. The influence of confinement
0
stiffness (Ecol), column friction angle ðϕcol Þ and potential upon the settlement performance of small groups of columns
increased coefficient of lateral earth pressure (K) arising from was investigated by analysing single columns, 2  2, 3  3,
installation were considered, as well as the presence of a stiff 4  4 groups and infinite column grids at various spacings,
crust typical of soft clay profiles and the position of columns which represent the spectrum of A/AC values typically
relative to the edge of the footing. These variables were encountered in practice, as identified in the McCabe et al.
investigated for groups of four, five and nine columns, a subset (2009) database (Fig. 2 and Table 1). Single columns and
of the full parametric study. infinite grids of columns provide a useful frame of reference

Fig. 2. Column configurations to examine influence of column confinement.

Table 1
Column configurations to investigate factors influencing column confinement.
0
No. of columns, N s B A/AC L E50,col ϕcol
(m) (m) (–) (m) (MPa) (1)

1 1, 1.5, 2 1, 1.5, 2 3.5, 8.0, 14.1 1.0–13.9 70 45


4 1, 1.5, 2 2, 3, 4 3.5, 8.0, 14.1 1.0–13.9 70 45
9 1, 1.5, 2 3, 4.5, 6 3.5, 8.0, 14.1 1.0–13.9 70 45
16 1, 1.5, 2 4, 6, 8 3.5, 8.0, 14.1 1.0–13.9 70 45

Fig. 3. Column configurations to examine influence of column position relative to footing edge.

Fig. 4. Column configurations to examine the influence of column compressibility, strength, installation effects and the presence of a stiff crust.
M.M. Killeen, B.A. McCabe / Soils and Foundations 54 (2014) 760–776 763

Table 2
Column configurations to investigate influence of other parameters.
0
No. of columns, N s B A/AC L E50,col ϕcol
(m) (m) (–) (m) (MPa) (1)

4 1.0 3 8.0 1.0–13.9 70 45


4 1.5 3 8.0 1.0–13.9 70 45
4 2.0 3 8.0 1.0–13.9 70 45

4 2.0 3 8.0 1.0–13.9 30, 50, 70 45


5 1.0 3 6.4 1.0–13.9 30, 50, 70 45
9 1.0 3 3.5 1.0–13.9 30, 50, 70 45

4 2.0 3 8.0 1.0–13.9 70 40, 45, 50


5 1.0 3 6.4 1.0–13.9 70 40, 45, 50
9 1.0 3 3.5 1.0–13.9 70 40, 45, 50

for the other configurations as they correspond to zero and


full lateral confinement, respectively. Columns are posi- CL
tioned on a square grid and the edge of footings is located at
a distance of half the column spacing (0.5  s) from the 8B

centre-line of the outer row of columns. CL


½B

2.2.2. Influence of other parameters


The influence of column position relative to the footing
edge was investigated by examining the settlement perfor-
mance of a group of four columns which are progressively
moved towards the footing edge beneath a 3 m square
footing (Fig. 3).
The influence of column stiffness, friction angle, coeffi-
cient of lateral earth pressure and the presence of the stiff
crust was investigated for three configurations of columns
illustrated in Fig. 4. The influence of column stiffness was
investigated by varying E50 (E50 ¼ secant Young's modulus
Fig. 5. Typical mesh adopted for FEA employing symmetry and with locally
at 50% of the deviator stress) from 30 to 70 MPa, while the
refined mesh around footing.
friction angle was varied from 401 to 501 in accordance with
the experience of McCabe et al. (2009) for the bottom-feed
system.
The changes to the in situ stress regime due to the
installation of stone columns are difficult to quantify due to 2.3. Pad footing
large strains and vibrations associated with poker penetra-
tion. Many authors (Kirsch, 2006; Castro, 2007; Gäb et al., Pad footings of 0.6 m thickness and founded at 0.6 m below
2007) have reported an increase in pore pressure, horizontal ground level (Fig. 1) are modelled as a linear elastic material
stress and stiffness in the surrounding soil. However, the with properties representative of concrete (i.e. unit weight,
increase in horizontal stress and soil stiffness is negated by γ¼ 24 kN/m3; Young's modulus, E¼ 3  107 kPa and Poisson's
remoulding and dynamic excitation of the soil close the ratio, υ¼ 0.15). The pad footings act as a rigid loading cap on
column. Despite this complexity, Elshazly et al. (2008a) top of the stone columns. Since the settlement performance of
attempt to capture column installation effects by increasing stone columns is usually the governing criterion for the design
the coefficient of lateral earth pressure (K0) in the surround- of foundations on soft soils (e.g. Priebe, 1976) and is also
ing soil. The authors conducted a series of axisymmetric the focus of this paper, a typical working pressure of 50 kPa
FEA to back-analysis field load tests described by Mitchell (i.e. within the range quoted by Serridge and Sarsby, 2008) has
and Huber (1985) and suggest conservative estimates of K0 been applied to each footing; this pressure corresponds to a
in the range 0.85–1.7, which increase at lower area ratios factor of safety of approximately 2.75 on the ultimate bearing
(A/AC ¼ 2.5–4.8). capacity of an unreinforced footing at the Bothkennar site
A summary of the parameters varied is outlined in Table 2. (Jardine et al., 1995).
764 M.M. Killeen, B.A. McCabe / Soils and Foundations 54 (2014) 760–776

2.4. Meshing and boundaries 2.6. Drainage conditions

The FE mesh comprised 15-noded wedge isoparametric PLAXIS 3D enables long-term behaviour to be modelled in
elements and the necessary detailed mesh sensitivity studies two ways: (i) drained analysis, using effective stress para-
are reported in Killeen (2012). The lateral boundaries of the meters and (ii) undrained analysis, using effective stress
computational domain allow vertical movement but restrict parameters, followed by consolidation. The use of effective
radial movement, and boundary sensitivity studies indicated stress parameters allows the increase of shear strength with
that these should be positioned at a distance of 8B from the effective stress to be captured. Method (ii) is a closer
centre of the pad to avoid influencing the results. The bottom simulation of the process in reality as a large bulk modulus
boundary is influenced by the ground profile which is is added to the soil stiffness matrix in order to mimic undrained
discussed in Section 4. In general, it was possible to exploit conditions, and subsequently removed during the consolidation
symmetry to save computation time, as illustrated by the mesh phase to allow the transfer of stress to soil particles. However,
in Fig. 5. runs based upon method (ii) are far more time-consuming and
method (i) was preferred in this research. Fig. 6 shows that for
a square footing with B ¼ 3 m, loaded to 50 kPa and supported
2.5. Column–soil interface by various configurations of stone columns outlined in the
inset to the figure, approaches (i) and (ii) produce quite similar
The column compaction process involves forcing the stone (if not perfectly consistent) results.
laterally into the host soil. Since the stone becomes tightly
interlocked with the surrounding soil, no discrete interface
zone exists as might be expected for a pile. Perfect adhesion is
assumed at the column–soil interface and, in keeping with 3. Hardening Soil model
standard practice for modelling stone columns, interface
elements are not used (e.g. Kirsch, 2006; Gäb et al., 2008; Both the stone columns and the host soft clay are modelled
Shahu and Reddy, 2011). Many authors report the develop- using the elasto-plastic Hardening Soil (HS) model. An
ment of a smear zone along the column–soil interface due to extension of the hyperbolic model developed by Duncan and
column installation (Han and Ye, 2001; Weber et al., 2008; Chang (1970), the HS model includes soil dilatancy, captures
Indraratna et al., 2013). This smear zone is characterised by unload–reload behaviour and accounts for shear and volu-
low permeability, which adversely affects the drainage capa- metric hardening. Schanz et al. (1999) advocate its suitability
city of stone columns and reduces the rate of consolidation. for both soft and stiff soils.
However, since this paper focuses on the long term (i.e. The basis of the HS model is that the stress–strain curve for
drained) settlement performance of stone columns, the con- a standard drained triaxial test may be approximated by a
solidation rate is not relevant. Other than field trials by Kirsch hyperbola, which is defined by a secant Young's modulus (E50)
(2006), there is no information on the extent to which column and an unload–reload modulus (Eur). The Mohr–Coulomb
installation increases lateral earth pressures in the surrounding failure criterion defines yielding and plasticity is accounted
soil, but increased K0 (coefficient of lateral earth pressure for using stress-dilatancy theory developed by Rowe (1962).
at rest) values are considered as a variable in the parametric The stiffness of soils (E) tends to increase with confining stress
study. and this is captured by the HS model using a power law:
 m
E σ
¼ ð3Þ
Settlement improvement factor, n
Eref pref
1.0 1.5 2.0 2.5 3.0 3.5 4.0
0 where σ is the confining stress and Eref is the reference soil
4 cols - s = 2.0 m stiffness for a given confining pressure, pref. The exponent, m,
5 cols - s = 1.0 m
2 9 cols - s = 1.0 m controls the relationship between soil stiffness and confining
pressure and Brinkgreve and Broere (2006) propose a value of
Column length, L (m)

Drained
4 Undrained & 1.0 for soft soils.
Consolidation
6

10

12

14

Fig. 6. Influence of analysis type upon settlement improvement factors for


various groups of columns. Fig. 7. Bothkennar test site location (Nash et al., 1992b).
M.M. Killeen, B.A. McCabe / Soils and Foundations 54 (2014) 760–776 765

Table 3
Summary of material parameters adopted for Bothkennar test site.

Soil parameter Crust Upper Carse clay Lower Carse clay

Depth (m) 0.0–1.5 1.5–2.5 2.5–14.5


Bulk unit weight, γ (kN/m3) 18.0 16.5 16.5
Over-consolidation ratio – – 1.5
Pre-overburden stress (kPa) 15 15 –
Coefficient of lateral earth pressure, K0 1.5 1.0 0.75
Effective cohesion, c' (kPa) 3 1 1
Angle of internal friction, ϕ' (1) 34 34 34
Initial voids ratio, e0 1.0 1.2 2.0
Compression index, CC 0.07 0.25 1.12
Swelling index, CS 0.01 0.03 0.16
Reference pressure, pref (kPa) 13 20 30
Vertical coefficient of permeability, kvert (m/day) 6.9  10  5 6.9  10  5 6.9  10  5
Horizontal coefficient of permeability, khorz (m/day) 1.0  10  4 1.0  10  4 1.0  10  4

The HS model does not account for the effects of secondary indicates that other post-depositional processes such as ageing
settlement and, consequently, this paper focuses on the long have occurred at the Bothkennar test site. Nash et al. (1992b)
term primary settlement of small groups of stone columns. measured the in situ coefficient of earth pressure at rest (K0)
and report a variation from 1.5 in the upper layers to 0.75
4. Development of soil profile and material properties at depth.
for HS model Allman and Atkinson (1992) conducted a series of triaxial
tests to determine the strength characteristics of reconstituted
4.1. Bothkennar test site Carse clay. The Carse clay was formed into a slurry (at a water
content of 1.25 times the liquid limit) and then one-
The Bothkennar test site is located along the Firth of Forth, dimensionally compressed into normally-consolidated or (with
near Grangemouth in Scotland (Fig. 7). The site is located swelling) into slightly over-consolidated states before shearing.
in a historic tidal estuary and primarily consists of soft uni- A critical state friction angle of ϕ' ¼ 341 is reported, which is
form clay, which is extensively characterised in Géotechnique larger than would be expected for a high plasticity clay but
(1992). The clay, often referred to as Carse clay, rests upon attributable to the high proportion of angular silt particles. In
Bothkennar gravel and is overlain by a stiff crust. The Carse order to ensure numerical stability, nominal values of 3 kPa
clay varies in thickness from 13 to 19 m across the site and is and 1 kPa were adopted for the effective cohesion (c0 ) in the
classified by Paul et al. (1992) as a silty clay with an organic stiff crust and Carse clay layers, respectively.
content of 3–8%. Piezometers indicate that the groundwater The one-dimensional stiffness properties of the Carse clay
conditions are hydrostatic and the water table varies from 0.5 were investigated by Nash et al. (1992a) through a series of
to 1.0 m below ground level. incremental load, restricted flow and constant rate of strain
The soil profile for the subsequent FEA is divided into three tests. The initial voids ratio (e0) and coefficient of compression
layers: crust; upper Carse clay and lower Carse clay. The (CC) are reported for various depths and these can be readily
development of material properties, outlined in Table 3, for converted into oedometric moduli (Eoed) for the HS model
each of these layers is described in the remainder of this using Eq. (4) (Brinkgreve and Broere, 2006):
section.
Hight et al. (1992) suggest that the Bothkennar test site may 2:3ð1 þ e0 Þpref
have been subjected to various post-depositional processes oed ¼
Eref ð4Þ
CC
such as erosion, changes in groundwater level and bonding.
These processes correspond to a 15 kPa drop in effective The relationship between E50 and Eoed is defined by:
overburden stress, which agrees favourably with yield stress
ratios (equivalent to over consolidation ratios, OCR) measured E50 ¼ 1:25E oed ð5Þ
in the upper layers by Nash et al. (1992a) in a series of
oedometer tests. In this paper, the stress history of the upper While the coefficient of swelling (CS) was not routinely
layers is modelled using a single pre-overburden pressure measured by Nash et al. (1992a), Allman and Atkinson (1992)
(POP) value of 15 kPa (POP is defined as the difference found the slopes of the normal compression (λ) and swelling
between the maximum vertical effective stress experienced (κ) lines to be 0.181 and 0.025, respectively. This ratio was
0
previously ðσ max Þ and the current in situ vertical effective stress used to determine the coefficient of swelling, i.e. λ/κ ¼ CC/
0 0 0
ðσ 0 Þ). However, a higher stress state ðOCR ¼ σ max = σ 0 ¼ 1:5Þ CS ¼ 7.2. The coefficient of swelling can be converted into an
is more appropriate for depths greater than 4 m, which unload–reload modulus using Eq. (6) (Brinkgreve and Broere,
766 M.M. Killeen, B.A. McCabe / Soils and Foundations 54 (2014) 760–776

2006): 17
Plaxis 3D Foundation
15 Priebe (1995): n0

Settlement improvement factor, n


2:3ð1 þ e0 Þð1þ νÞð1 2υÞpref
Priebe (1995): n2

ur ¼
E ref ð6Þ Pulko et al. (2011)
13
C S ð1  νÞ Balaam & Booker (1981)
11

9
4.2. Stone backfill
7

A bulk unit weight of γ ¼ 19 kN/m is adopted for the stone 3 5

backfill which is similar to values adopted by Mitchell and 3


Huber (1985), Gäb et al. (2008) and Domingues et al. (2007a). 1
An angle of internal friction of ϕ' ¼ 451 is adopted as the 2 4 6 8 10 12 14 16
default for the stone backfill on the basis of the finding that Area ratio, A/Ac

ϕ' ¼ 401 is conservative for stone columns formed using the Fig. 9. Comparison of settlement improvement factors for an infinite grid of
modern dry bottom feed system (McCabe et al., 2009). Kirsch end-bearing columns with analytical design methods.
(2004) proposed similar values. The angle of dilatancy is
related to the angle of internal friction through the well-known
empirical relationship, ψ0 ¼ ϕ0 –301. A nominal value of effec-
exceeded 8 mm/hr. Pad A was founded at 0.8 m below ground
tive cohesion (c0 ¼ 1 kPa) is adopted to ensure numerical
level and was 2.2 m wide. A second pad (Pad B) was not
stability.
modelled as significant creep settlements are incorporated in
Barksdale and Bachus (1983) observed that Young's moduli
the settlements recorded over a period in excess of 2 years.
for stone columns, which were back-calculated from measured
The load test on Pad A was simulated with PLAXIS 3D
settlements and recommended by other authors, vary from 30
Foundation as an undrained loading (with effective stress
to 58 MPa. However, since columns installed using the dry
material parameters) due to the short duration of the load test.
bottom feed system have been shown by McCabe et al. (2009)
However, the crust in the adopted soil profile is modelled as a
to exhibit better settlement performance than other systems, it
drained material as the lower part of the crust contains a
was deemed appropriate to adopt a higher Young's modulus
significant proportion of shelly fragments (Nash et al., 1992b)
for the stone backfill. Default values of E50 = 70 MPa and
and the upper part of the crust is above the groundwater level.
Eur ¼ 210 MPa were adopted. The exponent, m, which controls
In Fig. 8, it is clear that PLAXIS 3D Foundation captures
the stress dependency of soil stiffness for the HS model was
the trend of settlement versus applied pressure quite well,
chosen as m ¼ 0.3. These stiffness parameters (i.e. E50, Eur and
despite slightly over-predicting the settlement at low working
m) are in keeping with those adopted by Gäb et al. (2008).
loads. This may be explained as the location where the load
tests were conducted by Jardine et al. (1995) had a localised
4.3. Validation of Bothkennar soil profile shelly layer 0.3 m in thickness beneath the base of the footing,
which would have enhanced the settlement performance of the
Instrumented field load tests on unreinforced rigid pad footing.
footings at the Bothkennar test site, documented by Jardine
et al. (1995), were modelled as an exercise in validating the HS 5. Comparison of PLAXIS 3D and analytical design
parameters adopted for the soil. Pad A was loaded to failure method output
using kentledge in less than five days, with pauses in loading
occurring overnight and whenever the rate of settlement The settlement performance of an infinite grid of end-
bearing stone columns from PLAXIS 3D Foundation is
compared with various analytical design methods in Fig. 9.
140
The settlement performance is evaluated using a settlement
120 improvement factor, n, which is defined as the ratio of the
Jardine et al. (1995) PLAXIS 3D Foundation
Applied pressure (kPa)

100
settlement of an untreated footing (i.e. no stone columns) to
that of a treated footing (i.e. with columns). Settlement
80
improvement factors in Fig. 9 correspond to the application
60 of a 50 kPa uniform pressure on a concrete footing. A brief
summary of the design methods chosen for comparison,
40
adapted from Sexton et al. (2013), is presented in Table 4.
20 Another design method worthy of note (which is not listed
0 in Table 4) is that proposed by Castro and Sagaseta (2009).
0 20 40 60 80 100 120 140 160 180 200 The authors develop a comprehensive closed-form solution
Vertical displacement (mm) which accounts for the effects of both radial consolidation
Fig. 8. Comparison of PLAXIS 3D Foundation with actual load–displacement and loading history upon the settlement of stone columns.
behaviour for a pad footing observed by Jardine et al. (1995). However, for monotonic loading, as is the case with the FEA
M.M. Killeen, B.A. McCabe / Soils and Foundations 54 (2014) 760–776 767

Table 4
Summary of analytical design methods chosen for comparison.

Author Column behaviour Column yielding

Balaam and Booker (1981)  Elastic Not considered


 Accounts for radial expansion due to bulging
Priebe (1995)  Rigid-plastic and incompressible Mohr–Coulomb
 Modification factors introduced to account for:
(i) Column compressibility
(ii) Overburden stress

Pulko et al. (2011)  Elastic–plastic with a non-associated flow rule based on a constant angle of dilation Mohr–Coulomb
 Accounts for radial expansion of column and initial stress state in surrounding soil.

in this paper, the solution proposed by Castro and Sagaseta Pulko et al. (2011) is an extension of Balaam and Booker
(2009) is similar to Pulko et al. (2011) at the final state. (1981), as it accounts for column yielding, thereby predicting
All of the aforementioned design methods are based on the more realistic n values. It appears that columns are in an elastic
unit cell concept which is implemented in the FEA. However, state at low area ratios as Pulko et al. (2011) and Balaam and
the design methods are based upon more simplistic material Booker (1981) predict similar n values. However, as area ratio
models and it is necessary to select appropriate material increases the extent of plasticity becomes more pronounced
parameters which are consistent with the FEA: and Pulko et al. (2011) predict lower n values which agree
favourably with PLAXIS 3D Foundation.
 Priebe (1995) and Pulko et al. (2011) both account for The basic (n0) and modified (n2) settlement improvement
the initial stress state of the surrounding soil using the factors developed by Priebe (1995) are also shown in Fig. 9.
coefficient of lateral earth pressure at rest, K0. While Pulko The basic settlement improvement factors under-predict n
et al. (2011) allow for a direct input of K0, Priebe (1995) values, whereas the modified settlement improvement factors,
assumes that K0 ¼ 1 as a result of column installation. which account for column compressibility and overburden
Consequently, it was necessary to modify Priebe (1995) to stress, are in closer agreement with PLAXIS 3D Foundation. It
allow for alternative values of K0 for the soil for compara- can be seen that Priebe (1995) under-estimates n values at low
tive purposes. area ratios. This is due in part to the assumption of rigid-plastic
 While the HS model requires the selection of two material behaviour for the columns, which neglects strain developed
parameters (E50 and Eur) to capture the stress path depen- when columns are in an elastic state. The effect of this
dency of soil stiffness, the design methods are based on a assumption diminishes with increasing area ratio, as the extent
single stiffness modulus. E50 is adopted for the stone of plasticity becomes more pronounced and plastic strains
column as this is suitable for primary loading (Brinkgreve dominate the settlement of stone columns. At high area ratios,
and Broere, 2006). However, a higher stiffness modulus is n2 values slightly over-estimate PLAXIS 3D Foundation. This
adopted for the surrounding soil, as the soil is subject to may be explained as Priebe's modification for column com-
lower stress levels (due to stress concentrations which pressibility is based upon the constrained modular ratio of the
develop in the stiffer column) and the fact that soils tend stone column and soil (i.e. oedometric conditions). This
to exhibit increasing stiffness at low strain levels. An assumption over-estimates n values as it does not account for
average of E50 and Eur is chosen to represent the stiffness column bulging.
modulus of the surrounding soil. This exercise gives early confidence in the ability of
 The stiffness of soil tends to increase with confining PLAXIS 3D Foundation to capture stone column behaviour.
pressure and this is captured by the HS model using the
power law outlined in Eq. (3). The stress level dependency 6. Results of FEA
of soil stiffness is accounted for in the analytical design
methods by averaging the minor principal stresses, and 6.1. Column confinement
subsequently determining the average stiffness, in each
layer. The minor principal stress is adopted for the confin- The effect of column configuration was examined as
ing stress in accordance with Brinkgreve and Broere (2006), outlined in Section 2.2. Various column lengths were
as a FEA by Balaam and Booker (1985) shows that an considered which allows the relationship between column
infinite grid of columns are in a triaxial stress state during length, area ratio and the number of columns to be
loading. investigated.

6.1.1. Infinite grid of columns


It can be seen in Fig. 9 that Balaam and Booker (1981) over- The influence of column length and area ratio upon the
estimate n values, which is due to the assumption of linear settlement performance of an infinite grid of columns is shown
elasticity for the stone column (Castro and Sagaseta, 2009). in Fig. 10. A negligible reduction in footing settlement is
768 M.M. Killeen, B.A. McCabe / Soils and Foundations 54 (2014) 760–776

Footing settlement (mm) Settlement improvement factor, n


0 200 400 600 1 3 5 7 9 11
0 0
Unit Cell
A/Ac = 3.5
2 2
A/Ac = 8.0
Column length, L (m)

Column length, L (m)


A/Ac = 14.1
4 4

6 6

8 8

10 10
Unit Cell
12 A/Ac = 3.5 12
A/Ac = 8.0
A/Ac = 14.1
14 14
(i) (ii)
Fig. 10. Variation of (i) footing settlement and (ii) settlement improvement factor with column length for an infinite grid of columns.

Settlement improvement factor, n Settlement improvement factor, n Settlement improvement factor, n


1 3 5 7 9 11 1.0 1.5 2.0 2.5 3.0 1.0 1.2 1.4 1.6 1.8
0 0 0
A/A = 3.5 A/A = 8.0 A/A = 14.1
1 col 1 col 1 col
2 2 2
2×2 2×2 2×2

Column length, L (m)


Column length, L (m)

Column length, L (m)

3×3 3×3 3×3


4 4×4 4 4×4 4 4×4
Unit cell Unit cell Unit cell
6 6 6

8 8 8

10 10 10

12 12 12

14 14 14

Settlement improvement factor, n Settlement improvement factor, n Settlement improvement factor, n


1.0 2.0 3.0 4.0 5.0 6.0 1.0 1.3 1.6 1.9 2.2 2.5 1.0 1.1 1.2 1.3 1.4 1.5
0 0 0
Normalised column length, L/B
Normalised column length, L/B

Normalised column length, L/B

2 2 2

4 4 4

6 6 6

8 8 8

10 A/A = 3.5 10 A/A = 8.0 10 A/A = 14.1


1 col 1 col 1 col
2×2 2×2 12 2×2
12 12
3×3 3×3 3×3
4×4 4×4 4×4
14 14 14

Fig. 11. Variation of settlement improvement factor with (a) column length and (b) normalised column length for small groups of columns spaced at area ratios of
(i) 3.5, (ii) 8.0 and (iii) 14.1.

observed in Fig. 10(i) with increasing column length up to 6.1.2. Small groups of columns
L E2 m. This is attributed to the stiff crust, which extends to The relative settlement performances of the various stone
1.9 m, as installing stone columns in an already competent column group configurations identified in Table 1 are shown in
layer will do little to enhance the settlement performance. a plot of n against column length (L) in Fig. 11(a). As with an
A significant reduction in footing settlement, which yields a infinite grid of columns, a negligible increase in n values is
corresponding increase in n values (Fig. 10(ii)), is observed observed for all columns shorter than L E2 m. However, n
with increasing column length thereafter in the Carse clay. values increase with column length thereafter and it appears
It appears that the option of using of longer columns will be that the influences of area ratio and column length upon the
most advantageous at low area ratios. settlement performance of small groups of stone columns are
M.M. Killeen, B.A. McCabe / Soils and Foundations 54 (2014) 760–776 769

somewhat coupled, as the influence of column length becomes columns, as the combined shaft and base resistance surpasses
more pronounced at low area ratios. the lateral confining stress with increasing column length.
A continuous increase in n values with column length is Columns which are bulging cannot transfer the applied load to
observed in Fig. 11(a–i) for columns at low area ratios depth and, thus, do not yield a significant increase in n values
(A/AC ¼ 3.5). An abrupt increase in n values is observed when with column length.
moving from floating to end-bearing columns (i.e. from The deformed shape of a 3  3 group of 8 m long columns
L ¼ 13 m to L ¼ 13.9 m). This suggests that closely-spaced at A/AC ¼ 3.5, 8.0 and 14.1 is illustrated through plots of
columns transfer a large proportion of the applied load to the total shear strain in Figs. 12(i), 12(ii) and 12(iii), respec-
base of columns. Furthermore, the magnitude of the abrupt tively. It can be seen in Fig. 12(i) that limited bulging occurs
increase in n values increases with the number of columns, in the group of closely-spaced columns, which reflects the
which indicates that the proportion of the applied load enhanced lateral restraint experienced by columns that
transferred to the base of columns increases for larger groups precipitates a punching mode of deformation. In contrast,
of columns. This is akin to a punching mode of deformation columns at higher area ratios tend to bulge in the upper
and is consistent with findings by Muir Wood et al. (2000) and layers and bend outwards from the central columns (Fig. 12
Black et al. (2011) who conducted a series of small scale (iii)). Punching is reduced at higher area ratios as column
laboratory studies on small groups of stone columns. Muir bulging minimises the magnitude of applied load transferred
Wood et al. (2000) observed four modes of deformation to the base of columns.
(bulging, shearing, punching and bending) and noted that The increased level of lateral confinement associated with
punching was most pronounced for short columns at low area larger groups of columns is evident in Fig. 11 as an increase in
ratios. Similar findings are reported by Black et al. (2011) who the number of columns (for a given area ratio) yields higher
observed the development of ‘block failure’ for a small group n values. However, this trend of increasing n values with group
of long, closely-spaced columns (A/AC ¼ 2.5), which acted as a size does not hold for an infinite grid of floating columns
single entity and punched into the underlying soil. shorter than L E 7 m. This can be attributed to the constant
A constant increase in n values with column length is increment of vertical stress with depth associated with an
evident up to approximately L E 5 m for columns at A/AC ¼ 8.0 infinite grid of columns, which induces significant settlement
and 14.1 in Fig. 11(a–ii) and (a–iii), respectively. However, the beneath the base of floating columns. The increase in n values
rate of increase in n values with column length slows down with column length is most pronounced for an infinite grid of
thereafter, which suggests that the mode of deformation may columns at all area ratios and the positive effects of confine-
have transitted from punching to bulging at higher area ratios. ment are evident for longer columns. The different variations
Similar observations were made by Hughes and Withers of n values with column length observed for small groups and
(1974), McKelvey et al. (2004) and Black et al. (2011) in infinite grids of columns highlights the difficulty of employing
small scale laboratory tests on single and small groups of settlement design methods based on the unit cell model to

(×10-3) (×10-3) (×10-3)

Low strain in
soil bounded by
outer columns Development
of shear zone

Uniform strain at
base of columns

(i) (ii) (iii)


Fig. 12. Distribution of total shear strains through a cross-section of a 3  3 group of 8 m long columns at A/AC of (i) 3.5, (ii) 8.0 and (iii) 14.1.
770 M.M. Killeen, B.A. McCabe / Soils and Foundations 54 (2014) 760–776

Settlement improvement factor, n Settlement improvement factor, n


1.0 1.2 1.4 1.6 1.8 1.0 2.0 3.0 4.0
0 0
4 cols - s = 1.0 m 4 cols - s = 2.0 m
4 cols - s = 1.5 m 5 cols - s = 1.0 m
2 4 cols - s = 2.0 m 2
9 cols - s = 1.0 m
Column length, L (m)

Column length, L (m)


Footing width, B = 3 m E50,col = 30 MPa
4 4 E50,col = 50 MPa
E50,col = 70 MPa
6 6

8 8

10 10

12 12

14 14
Fig. 13. Influence of column position upon settlement improvement factors. Fig. 14. Influence of column stiffness upon settlement improvement factors.

capture the settlement performance of small groups of stone


columns.
The same data are presented in Fig. 11(b), this time with the
column length normalised by the footing width (i.e. L/B); n
values are not presented for the ‘unit cell’ model as the footing
width is theoretically infinite. The classical Boussinesq (1885)
solution for the distribution of vertical stress in an elastic
medium indicates that the increment of stress applied at the
surface is considered negligible at a depth of 2B (B ¼ footing
width). Muir Wood et al. (2000) suggest that the dominant
strain in columns is dependent upon footing width and
postulate the existence of a critical length, beyond which the
mode of deformation changes from punching to bulging.
However, Fig. 11(b) shows that the settlement performance
of small groups of stone columns is dependent upon a
combination of area ratio, column length and number of
Fig. 15. Influence of column strength upon settlement improvement factors.
columns, rather than footing width alone. The laboratory tests
undertaken by Muir Wood et al. (2000) incorporate a weaker
0
sand column ðϕcol ¼ 301Þ, which is loaded to failure in a and thus closer to these shear zones, may also explain the
homogenous soil profile. All of these factors increase the enhanced settlement performance.
likelihood of bulging failure, which may have guided the
author's interpretation of a critical length. 6.2.2. Column stiffness
The influence of column stiffness upon the settlement perfor-
mance of various configurations of columns (see Fig. 4) is shown
6.2. Parametric study
in Fig. 14. As expected, an increase in n values is observed with
increasing column stiffness. This is consistent with the work of
6.2.1. Column position
Domingues et al. (2007a) who conducted a FEA on a large group
The influence of column position relative to the footing
of columns (A/AC ¼ 5.3) supporting an embankment. The increase
edge (see Fig. 3) upon n values is shown in Fig. 13. It
in n values with column stiffness is most pronounced for columns
appears that small benefits can be gained by positioning
at low area ratios, as the degree of plasticity within columns is
columns closer to the footing edge, confirming the merit of
reduced due to the enhanced column confinement. Consequently,
this strategy which is sometimes used in practice. This may
columns at low area ratios are more susceptible to changes in
be attributed to stress concentrations which develop beneath
elastic stiffness moduli.
the edges of rigid footings. Columns positioned closer to
this zone have the potential to absorb a larger proportion of
the applied load and, hence, develop higher n values. Wehr 6.2.3. Column friction angle
0
(2004) conducted a FEA on small groups of stone columns The influence of the angle of internal friction ðϕcol Þ upon the
and observed the development of shear zones which extend settlement performance of various configurations of stone
from the edges of footings to a depth beneath the centre of columns (see Fig. 4) is shown in Fig. 15. The angle of internal
footings. Positioning columns closer to the footing edge, friction controls the plastic deformation of columns, i.e.
M.M. Killeen, B.A. McCabe / Soils and Foundations 54 (2014) 760–776 771

0
deformation which occurs post-yielding. Increasing ϕcol from 6.2.5. Presence of stiff crust and lower Carse clay layer
the default value of 451 to 501 has the lowest influence on the The influence of the lower Carse clay and stiff crust upon
group of nine columns, as columns at low area ratios exhibit the the settlement performance of small groups of stone columns
0
lowest degree of plasticity. Decreasing ϕcol from the default (see Fig. 4) is analysed by adjusting the standard soil profile in
value of 45–401 is most pronounced for columns at low area two steps, as illustrated in Fig. 17. The importance of soil
ratios, as the groups of four and five columns (at higher area stiffness variation with depth is clearly evident in Fig. 18 as n
ratios) are already in an advanced state of plasticity. values are lowest for Profile 1, which contains the weak lower
Carse clay (not present in Profiles 2 and 3). A sharp increase in
n values occurs with increasing column length up to L ¼ 3 m
for Profile 3. However, the rate of increase in n values with
6.2.4. Coefficient of lateral earth pressure column length reduces thereafter. This is most pronounced for
Fig. 16 shows the influence of K0 for the lower Carse clay smaller groups of columns (i.e. higher area ratios) and suggests
layer upon the settlement performance of various configurations that a critical length may exist for columns formed in Profile 3.
of columns outlined in Fig. 4. On the basis of previous research The distribution of vertical strain along the length of a 2  2
outlined in Section 2.2, the in situ stress state of the upper Carse group of 4 m long columns is shown in Fig. 19. A comparison
clay and crust is deemed sufficiently high (K0 Z1.0) and has not of Profiles 1 and 2 with Profile 3 reveals that column bulging
been modified. An increase in K0 generally has a positive becomes shallower and more pronounced with the absence of a
influence on n values for all configurations of columns. The stiff crust. It can also be seen that the magnitude of vertical
influence of K0 upon n values is most pronounced for groups of strain in the underlying soil (i.e. below a depth of 4 m) is
four and five columns, as columns at high area ratios tend to higher for Profile 1 relative to Profiles 2 and 3, which indicates
exhibit bulging and, hence, are dependent upon the lateral the increased potential for column punching in weaker soil
support provided by the surrounding soil. In contrast, columns strata. Profile 3 is a homogenous soil profile and is similar to
at low area ratios transfer the applied load to the base of that used in laboratory studies. The low lateral support
columns by punching and are less dependent upon K0 of the provided by the upper layers, and subsequent bulging, may
surrounding soil. explain why authors (Muir Wood et al., 2000; McKelvey et al.,
2004; Black et al., 2011) have postulated critical lengths in the
Settlement improvement factor, n past. However, it can be seen that critical lengths are less well
1.0 2.0 3.0 4.0 5.0 defined for more realistic soft soil profiles which incorporate
0 stiff upper layers. Fig. 19 highlights the importance of the stiff
4 cols - s = 2.0 m
5 cols - s = 1.0 m
crust upon the settlement performance of stone columns. This
2
9 cols - s = 1.0 m is a salient feature of many soft soil profiles (e.g. Väsby and
Skå-Edeby test fill areas (Larsson, 1986); Ellingsrud test fill
Column length, L (m)

K0,soil = 0.75
4 K0,soil = 1.00
K0,soil = 1.25
area (Karlsrud and Haugen, 1979); and Murro test embank-
6 ment (Karstunen and Yin, 2010)) and should be taken into
account when designing stone columns in soft soils.
8

10
7. Development of a simplified design framework

12 It was shown in the previous section that column confine-


ment has a significant influence upon the settlement perfor-
14 mance of small groups of stone columns. The degree of
Fig. 16. Influence of the coefficient of lateral earth pressure upon settlement column confinement depends upon many factors such as area
improvement factors. ratio, the number of columns and column length. The influence

Fig. 17. Soil profile adopted to investigate influence of lower Carse clay (Profile 2) and stiff crust (Profile 3).
772 M.M. Killeen, B.A. McCabe / Soils and Foundations 54 (2014) 760–776

of the number of columns upon the settlement performance of Settlement improvement factor, n
small groups of stone columns is examined using a settlement 1.0 1.5 2.0 2.5 3.0 3.5
ratio which is defined as the ratio of the settlement of a small 0 Profile: 1 2 3

group (s) to that of an infinite grid (suc) of columns. Crust


UCC
Crust
UCC
UCC
2 LCC

Column length, L (m)


4
7.1. Previous research using settlement ratios removing Crust
6 removing LCC
7.1.1. Numerical methods
Elshazly et al. (2008b) conducted a series of axisymmetric 8
FEA to examine the settlement performance of small groups of
10
end-bearing stone columns. The authors use s/suc to examine
4 cols - 2.00 m
the trade-off between the decay of vertical stress with depth 12 Profile 1
Profile 2
and loss of lateral confinement associated with small groups of Profile 3
stone columns. For small groups of stone columns, the decay 14
of vertical stress with depth dominates due to the small footing
widths and, consequently, s/suc o 1. However, as the number Settlement improvement factor, n
of columns increases, the distribution of vertical stress with 1.0 1.5 2.0 2.5 3.0 3.5
0
depth beneath footings becomes closer to that for unit cell Profile: 1 2 3
Crust Crust UCC
conditions and the influence of the loss of lateral confinement 2 UCC UCC
LCC
becomes more prevalent. This yields s/suc 41 which indicates

Column length, L (m)


that settlement design methods based on the unit cell concept 4
are unconservative. These ratios are difficult to justify; the high removing Crust
6
s/suc values may be due in part to the flexible loading condition
and weak surrounding soil (i.e. low lateral support). 8
removing LCC

10
7.1.2. Analytical methods 5 cols - 1.00 m
Priebe (1995) presents a design chart which relates s/suc to 12 Profile 1
Profile 2
normalised column length (L/d). The s/suc values are devel- Profile 3
oped from numerous calculations which consider the reduction 14
of vertical stress with depth beneath small loaded areas and, Settlement improvement factor, n
also, a lower bearing capacity for the outer row of columns in a 1.0 2.0 3.0 4.0 5.0
group. No direct reference is made to footing area (A) in the 0 Profile: 1 2 3
design chart and it appears that for a given number of columns Crust Crust UCC
UCC UCC
s/suc does not depend upon area ratio. However, an increase in 2 LCC

area ratio results in higher settlements for small groups and


Column length, L (m)

4
infinite grids of columns. Priebe (1995) claims that the increase
in ‘s’ and ‘suc’ compensate such that s/suc is acceptable for 6 removing
A/AC o 10. Crust
8
removing LCC
7.2. Comparison of s/suc from Priebe (1995) and PLAXIS 3D 10
foundation for end-bearing columns 9 cols - 1.00 m
12 Profile 1
Profile 2
Settlement ratios determined from PLAXIS 3D Foundation Profile 3
14
and those of Priebe (1995) are shown in Fig. 20. It can be seen
Fig. 18. Influence of the lower Carse clay and the stiff crust upon settlement
that s/suc increases for larger groups of columns. An increase in
improvement factors for (a) four, (b) five and (c) nine columns beneath a 3 m
the number of columns leads to higher levels of lateral square footing.
confinement and also higher levels of vertical stress with
depth (due to larger footing widths). Therefore, the boundary
conditions approach unit cell conditions with an increasing 7.3. Settlement ratios for floating stone columns
number of columns and s/suc tends towards unity. It can also
be seen that s/suc increases with area ratio. The design curve Currently no study has investigated the variation of s/suc for
developed by Priebe (1995) captures the variation of s/suc quite floating stone columns. The effect of floating columns is
well for A/AC ¼ 3.5, but diverges at high area ratios. This introduced into Fig. 21 using the L/H term, where H is the
confirms that Priebe's (1995) method is only applicable for low thickness of the soil deposit (H ¼ 13.9 m for the Bothkennar
area ratios and is unconservative for widely spaced columns. deposit in this study). The influence of area ratio upon s/suc for
M.M. Killeen, B.A. McCabe / Soils and Foundations 54 (2014) 760–776 773

floating columns is also shown in Fig. 20. For clarity, only Vertical strain, εy (10-3)
data specific to three column lengths are presented, namely: 0 10 20 30
L/H ¼ 0.22, 0.58 and 1.00, which correspond to L ¼ 3, 8 and 0 Bulging becomes shallower
13.9 m, respectively. 1 and more pronounced with

Depth beneath footing base (m)


3
2 absence of stiff crust
2
For a given B/L, it can be seen that s/suc reduces with
COLUMN
decreasing L/H. A decrease in L/H corresponds to a thicker soil 4
2 + 3

deposit beneath the base of columns. The settlement of the soil SOIL
1 Increased punching in
beneath the base of columns is much higher for an infinite grid 6 weaker deposits at depth
of columns compared to small groups due to the constant
increment of vertical stress with depth. Therefore, a decrease in 8
L/H leads to higher suc and, consequently, lower s/suc. 10 Soil Profiles:
1 2 3
It appears that area ratios only influence s/suc for end- Crust Crust UCC 4 cols - 2.00 m
bearing columns (L/H ¼ 1.0), which is due to the large 12 UCC
LCC
UCC Profile 1
Profile 2
reduction in suc for an infinite grid of long closely-spaced Profile 3
columns (Fig. 10(i)). In contrast, it can be seen that area ratios
do not influence s/suc for floating columns. For a given L/H, it Fig. 19. Distribution of vertical strain within and directly beneath 4 m long
columns, which are part of a 2  2 group in Profiles 1, 2 and 3.
appears that s/suc does not depend on the configuration of
columns and is directly proportional to the footing width (i.e. B/L).
The variation of s/suc for additional column lengths is shown 0.5

in Fig. 22. A somewhat linear relationship appears to exist


between s/suc and B/L for floating stone columns. However, 0.4

Settlement ratio, s/suc


care must be exercised in assuming a linear relationship for
0.3
end-bearing columns as s/suc from Priebe (1995) approach
unity somewhat asymptotically at high B/L. The linear 0.2
Priebe (1995)

relationship between s/suc and B/L for floating stone columns


is justified by the high coefficients of determination (R2). 0.1
PLAXIS 3D Foundation
A/Ac=3.5
Assuming a linear relationship between s/suc and B/L, the A/Ac=8.0
A/Ac=14.1
generalised relationship may be defined by: 0.0
  0 5 10 15 20
s B
¼α ð7Þ No. of columns
suc L
Fig. 20. Comparison of s/suc for groups of end-bearing columns from PLAXIS
The variation of α with L/H is shown in Fig. 23 and can be 3D Foundation with Priebe (1995).
captured by a quadratic equation:
 2   1.0
L L PLAXIS 3D Foundation
α ¼ 0:61 þ 0:1 þ 0:06 ð8Þ A/Ac = 3.5
H H 0.8
A/Ac = 8.0
A/Ac = 14.1
Settlement ratio, s/suc

Priebe (1995)
Therefore, it is possible to relate s/suc to B/L and L/H by the A/Ac = 3.5
0.6
following equation:
"  2   #   L/H = 1.00
s L L B 0.4
¼ 0:61 þ 0:1 þ 0:06 U ð9Þ L/H = 0.58
L/H = 0.22
suc H H L
0.2

The relevant value of suc in Eq. (9) will depend on whether


the columns are floating or end-bearing. However, analy- 0.0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5
tical design methods have only been developed for the end- Normalised footing width, B/L
bearing scenario (suc,eb). Therefore, the dependence of the
ratio suc,eb/suc on L/H has been investigated from the FEA, Fig. 21. Variation of s/suc with normalised footing width and column length.
with a view to rendering Eq. (9) amenable for practical use.
Suitable curve-fitted relationships are shown in Fig. 24 in the
form of Eq. (10): The settlement (s) of a group of stone columns can then be
  determined by combining Eqs. (9)–(11) as follows:
L suc; eb
¼ β ln þ1 ð10Þ
H suc
s ½0:61ðL=HÞ2 þ 0:1ðL=HÞþ 0:06 UðB=LÞ
¼ ð12Þ
The coefficients β are shown in Fig. 25 to vary approxi- suc; eb e8:71ððL=HÞ  1ÞðAC =AÞ
mately linearly with A/AC, as given by the following equation:
  It is recommended that either Pulko et al. (2011) or Castro
A
β ¼ 0:1148 ð11Þ and Sagaseta (2009), rather than Priebe (1995), is used as the
AC basis of estimating suc,eb in Eq. (12).
774 M.M. Killeen, B.A. McCabe / Soils and Foundations 54 (2014) 760–776

1.0 Area ratio, A/AC


0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0 16.0
Normalised settlement, s/suc

0.8 0.0

0.2
β = 0.1148(A/Ac)
0.6

β = (L/H - 1) / ln(suc,eb/suc)
0.4

0.4
0.6

0.8
0.2
1.0

1.2
0.0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 1.4
Normalised footing width, B/L
1.6
Fig. 22. Influence of normalised footing width and column length upon s/suc 1.8
for groups of columns.
Fig. 25. Relationship between β (which relates suc,eb/suc and L/H) with A/AC.

1.0
While the settlement of a large group of columns was not
measured at the same site, it was possible to estimate suc,eb ¼
0.8
47 mm using Pulko et al. (2011). This results in s/suc,eb ¼ 0.53,
α = 0.61(L/H) + 0.1(L/H) + 0.06
which is higher than s/suc,eb ¼ 0.29 determined using Eq. (12)
α = (s/suc) / (B/L)

R =1
0.6
(i.e. for L/H ¼ 0.45, B/L ¼ 0.33 and A/AC ¼ 3.58). The high
s/suc,eb values observed for Kirsch (2004) may be due in part to
0.4
undrained response of the surrounding soil as a result of the
short duration of the load test; the long term settlement of an
0.2
infinite grid (suc,eb) is much higher than a small group
of columns (s) as the soil is stressed to the full depth of the
0.0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 soil profile and, hence, will develop larger consolidation
Normalised column length, L/H settlement.
The availability of further data will allow the form of
Fig. 23. Relationship between α (which relates s/suc and B/L) with L/H.
Eq. (12) to be appraised more closely and modified as appro-
priate. However, this approach provides a useful framework
suc,eb / suc through which the behaviour of small groups of columns can
0.0 0.2 0.4 0.6 0.8 1.0 be distinguished from larger groups.
0.0
Normalissed column length, L/H

8. Conclusions
0.2

The settlement performance of various configurations of


0.4
columns beneath rigid square pad footings in the Bothkennar
soil profile is examined in this study.
0.6

8.1. Influence of column confinement


0.8

 The influences of area ratio, column length and the number


1.0
of columns upon the settlement performance of stone
Fig. 24. Relationship between suc,eb/suc with normalised column length (L/H). columns are somewhat inter-related, as the influence of
column length becomes more pronounced at low area ratios.
7.4. Comparison of simplified design framework  The unit cell model does not capture the reduction of
with field data vertical stress with depth beneath small loaded areas and,
hence, under-predicts n values for floating columns shorter
There is a shortage of field data with which the validity of than L≈7 m, corresponding to L/H ¼ 0.5 in this study.
Eq. (12) could be assessed – only one partially-suitable case
history is known to the authors. Kirsch (2004) reports on a
load test on a group of 5 columns beneath a 3 m square footing 8.2. Influence of other parameters
in Kuala Lumpur, Malaysia. Columns were installed to 9 m in
a 20 m deep soil deposit, which primarily consisted of soft clay  For a given number of columns, their position relative to the
underlain by sandy silt. The footing was loaded to 118 kPa in footing edge has a minor influence upon the settlement
under 12 days, and 25 mm settlement was observed at 50 kPa. performance.
M.M. Killeen, B.A. McCabe / Soils and Foundations 54 (2014) 760–776 775

 Increasing the column stiffness and strength increases Castro, J., Sagaseta, C., 2009. Consolidation around stone columns. Influence
n values. The influence of these parameters is related to of column deformation. Int. J. Numer. Anal. Methods Geomech. 33 (7),
851–877.
the degree of plasticity within columns, as increasing the
Domingues, T.S., Borges, J.L., Cardoso, A.S., 2007. Parametric study of stone
stiffness and strength is most significant at low and high columns in embankments on soft soils by finite element method. In: Proceedings
area ratios, respectively. of the 5th International Workshop on Applications of Computational Mechanics
 Accounting for column installation effects by increasing the in Geotechnical Engineering V, Guimaraes, pp. 281–291.
coefficient of lateral earth pressure at rest (K0) in the lower Duncan, J.M., Chang, C.Y., 1970. Nonlinear analysis of stress and strain in
Carse clay yields higher n values. The influence of K0 upon soil. ASCE J. Soil Mech. Found. Div. 96 (5), 1629–1653.
Elshazly, H., Elkasabgy, M., Elleboudy, A., 2008a. Effect of inter-column
the settlement performance of stone columns appears to be spacing on soil stresses due to vibro-installed stone columns: interesting
dependent upon the mode of deformation and is most findings. J. Geotech. Geol. Eng. 26 (2), 225–236.
pronounced at high area ratios, i.e. column bulging. Elshazly, H.A., Hafez, D.H., Mossaad, M.E., 2008b. Reliability of conven-
 The presence of the stiff crust has a significant influence tional settlement evaluation for circular foundations on stone columns. J.
upon the settlement performance of stone columns. The stiff Geotech. Geol. Eng. 26 (3), 323–334.
Gäb, M., Schweiger, H.F., Thurner, R., Adam, D., 2007. Field trial to
crust confines columns in the upper layers and forces investigate the performance of floating stone columns. In: Proceedings of
bulging to occur deeper, in weaker layers. Conclusions the 14th European Conference on Soil Mechanics and Geotechnical
about critical column length quoted in previous studies Engineering, Madrid, pp. 1311–1316.
appear to be specific to ultimate conditions and due to the Gäb, M., Schweiger, H.F., Kamrat-Pietraszewska, D., Karstunen, M., 2008.
Numerical analysis of a floating stone column foundation using different
unrealistic simplified ground profile modelled rather than
constitutive models. In: Proceedings of the 2nd International Workshop on
mechanisms particular to stone column behaviour. Geotechnics of Soft Soils, Glasgow, pp. 137–142.
Hight, D.W., Bond, A.J., Legge, J.D., 1992. Characterization of the Bothken-
nar clay: an overview. Géotechnique 42 (2), 303–347.
8.3. Simplified design framework Hughes, J.M., Withers, N.J., 1974. Reinforcing of soft cohesive soils with
stone columns. Ground Eng. 7 (3), 42–49.
Han, J., Ye, S.L., 2001. Simplified method for consolidation rate of stone
A simplified design framework is proposed in which the column reinforced foundations. J. Geotech. Geoenviron. Eng. ASCE
settlement (s) of either floating or end-bearing small groups 127 (7), 597–603.
can be estimated from the settlement of infinite grids of end Indraratna, B., Basack, S., Rujikiatkamjorn, C., 2013. Numerical solution of
bearing columns (suc,eb) and geometrical factors such as area stone column-improved soft soil considering arching, clogging, and smear
ratio (A/AC), normalised footing width (B/L) and normalised effects. J. Geotech. Geoenviron. Eng. 139 (3), 377–394.
Jardine, R.J., Lehane, B.M., Smith, P.R., Gildea, P.A., 1995. Vertical loading
column length (L/H). experiments on rigid pad foundations at Bothkennar. Géotechnique 45 (4),
573–597.
Acknowledgements Killeen, M.M., 2012. Numerical modelling of small groups of stone columns
(Ph.D. thesis). National University of Ireland, Galway.
Kirsch, F., 2004. Experimentelle und numerische Untersuchungen zum
The authors wish to acknowledge the support of the Irish
Tragverthalten von Rüttelstopfsäulengruppen (Dissertation). Technische
Research Council for Science, Engineering and Technology Universität Braunschweig.
EMBARK Initiative for funding the first author’s research. Kirsch, F., 2006. Vibro stone column installation and its effect on the ground
The authors are also extremely grateful for the input of Keller improvement. In: Proceedings of the International Conference on Numer-
Foundations (UK) and Mr.Brian Sexton (NUI Galway) to ical Simulation of Construction Processes in Geotechnical Engineering for
Urban Environment, Bochum, pp. 115–124.
this study.
Kirsch, F. 2008. Evaluation of ground improvement by groups of vibro stone
columns using field measurements and numerical analysis. In: Proceedings
References of the 2nd International Workshop on Geotechnics of Soft Soils, Glasgow,
pp. 241–248.
Allman, M.A., Atkinson, J.H., 1992. Mechanical properties of reconstituted Karlsrud, K., Haugen, T., 1979. Lang-tids belastningsforsøk på kvikkleire.
Bothkennar soil. Géotechnique 42 (2), 289–301. Nordic Geotechnical Meeting, 609–619 (Helsinki).
Balaam, N.P., Booker, J.R., 1981. Analysis of rigid rafts supported by granular Karstunen, M., Yin, Z.-Y., 2010. Modelling time-dependent behaviour of
piles. Int. J. Numer. Anal. Methods Geomech. 5, 379–403. Murro test embankment. Géotechnique 60 (10), 735–749.
Balaam, N.P., Booker, J.R., 1985. Effect of stone column yield on settlement Larsson, R., 1986.Consolidation of soft clays. Report No. 29, Swedish
of rigid foundations in stabilised clay. Int. J. Numer. Anal. Methods Geotechnical Institute, Linköping, 174p.
Geomech. 9, 331–351. McCabe, B.A., Nimmons, G.J., Egan, D., 2009. A review of field performance
Barksdale, R.D., Bachus, R.C., 1983. Design and Construction of Stone of stone columns in soft soils. Proc. Inst. Civil Eng. – Geotech. Eng. 162
Columns Report No. FHWA/RD-83/026. National Information Service, (6), 323–334.
Springfield, Virginia, USA. McKelvey, D., Sivakumar, V., Bell, A.L., Graham, J., 2004. Modelling
Black, J.A., Sivakumar, V., Bell, A., 2011. The settlement performance of vibrated stone columns in soft clay. Proc. Inst. Civil Eng. – Geotech.
stone column foundations. Géotechnique 61 (11), 909–922. Eng. 157 (3), 137–149.
Boussinesq, J., 1885. Applications des Potentiels à L'étude de L'équilibre et du Mitchell, J.K., Huber, T.R., 1985. Performance of a stone column foundation.
Mouvement des Solides Élastiques, Gauthier-Villars, Paris. J. Geotech. Eng., ASCE 11 (2), 205–223.
Brinkgreve, R.B.J., Broere, W., 2006. PLAXIS 3D Foundation Manual Muir Wood, D., Hu, W., Nash, D.F.T., 2000. Group effects in stone column
Version 2. PLAXIS BV. foundations: model tests. Géotechnique 50 (6), 689–698.
Castro, J., 2007. Pore pressures during stone column installation. In: Proceed- Nash, D.F.T., Sills, G.C., Davison, L.R., 1992a. One-dimensional consolida-
ings of the18th European Young Geotechnical Engineers' Conference, tion testing of the soft clay from Bothkennar. Géotechnique 42 (2),
Ancona. 241–256.
776 M.M. Killeen, B.A. McCabe / Soils and Foundations 54 (2014) 760–776

Nash, D.F.T., Powell, J.J.M., Lloyd, I.M., 1992b. Initial investigations of the deposit. In: Proceedings of the 2nd International Workshop on Geotechnics
soft clay test site at Bothkennar. Géotechnique 42 (2), 163–181. of Soft Soils, Glasgow, pp. 293–298.
Paul, M.A., Peacock, J.D., Wood, B.F., 1992. The engineering geology of the Sexton, B.G., McCabe, B.A., Castro, J., 2013. Appraising vibro settlement
Carse clay of the national soft clay research site, Bothkennar. Géotechnique prediction methods using the finite element method. Acta Geotech. (minor
42 (2), 183–198. corrections sought Jan 2013).
Priebe, H.J., 1976. Evaluation of the settlement reduction of a foundation Shahu, J.T., Reddy, Y.R., 2011. Clayey soil reinforced with stone column
improved by vibro-replacement. Bautechnik 5, 160–162. group: model tests and analyses. J. Geotech. Eng., ASCE 137 (12),
Priebe, H.J., 1995. The design of vibro replacement. Ground Eng. 28 (10), 1265–1274.
31–37. Sivakumar, V., Jeludine, D.K.N.M., Bell, A., Glynn, D.T., Mackinnon, P.,
Pulko, B., Majes, B., Logar, J., 2011. Geosynthetic-encased stone columns: 2011. The pressure distribution along stone columns in soft clay under
analytical calculation model. Geotext. Geomembr. 29 (1), 29–39. consolidation and foundation loading. Géotechnique 61 (7), 613–620.
Rowe, P.W., 1962. The stress-dilatancy relation for static equilibrium Wehr, W., 2004. Stone columns – single columns and group behavior. In:
of an assembly of particles in contact. Proc. R. Soc. A 269 (1339), Proceedings of the 5th International Conference on Ground Improvement
500–527. Techniques, Malaysia, pp. 329–340.
Schanz, T., Vermeer, P.A., Bonnier, P.G., 1999. The hardening-soil model: Weber, T.M., Springman, S.M., Gäb, M., Racansky, V., Schweiger, H.F.,
Formulation and verification. In: Brinkgreve, R.B.J. (Ed.), Beyond 2000 in 2008.Numerical modeling of stone columns in soft clay under an
Computational Geotechnics – 10 Years of PLAXIS International. embankment. In: Proceedings of the 2nd International Workshop on
A.A. Balkema, Rotterdam, Netherlands, pp. 281–290. Geotechnics of Soft Soils, Glasgow, pp. 305–311.
Serridge, C.J., Sarsby, R.W., 2008. A review of field trials investigating the
performance of partial depth vibro stone columns in a deep soft clay

You might also like