You are on page 1of 14

Small Stone-Column Groups: Mechanisms of Deformation at

Serviceability Limit State


Bryan A. McCabe1 and Micheál M. Killeen2

Abstract: Stone columns are a popular form of ground improvement that can be applied to a variety of foundation solutions involving soft soil
deposits. Many design methods are based on an assumption that columns with a length-to-diameter ratio greater than 4 bulge uniformly along
their length. This mode of deformation is not valid for small groups of columns in which interaction between individual columns and the footing
Downloaded from ascelibrary.org by Ryerson University on 10/06/16. Copyright ASCE. For personal use only; all rights reserved.

gives rise to more complex behavior. The deformational behavior of stone columns at serviceability limit state is of utmost importance, because
settlement rather than bearing capacity criteria are generally more onerous in soft soils. The results of some high-quality laboratory studies have
been informative in this respect, but the extrapolation of results from small-scale tests comes with the caveats associated with single-gravity
laboratory testing. The FEM, in combination with advanced constitutive models, can provide a link between the deformational behavior and
settlement performance of discrete column groups at working stress levels. A series of analyses was conducted to examine the influence of key
design parameters such as column length, spacing, and the number of columns. The influence of a stiff crust was also examined, because the
reference soil profile adopted for these analyses was that of the well-characterized Bothkennar test site. The following three modes of deforma-
tion were identified by examining the distribution of stress and strain within columns and the surrounding soil: individual punching, block
punching, and bulging. Individual and block punching developed in short or closely spaced columns, whereas block punching was more preva-
lent in larger groups. Bulging develops in long, widely spaced columns and generally occurs at the point of lowest lateral restraint. In addition,
novel parameters referred to as compression and punching ratios were developed to validate the in-depth analysis and identify the principal
design parameters that influence the behavior of small groups of stone columns. The results indicate that the mode of deformation is governed
by column spacing and length rather than the number of columns. This finding is important, because it indicates that settlement design methods
that are developed for large groups of columns are applicable to discrete groups at similar spacings and lengths. It was also observed that the
influence of the stiff crust forces bulging to occur deeper and might have led authors in the past to suggest the existence of a critical length.
DOI: 10.1061/(ASCE)GM.1943-5622.0000700. © 2016 American Society of Civil Engineers.
Author keywords: Footings; Stone columns; Deformational behavior.

Introduction hypothesis that forms the basis of many analytical design methods,
for both bearing capacity and settlement prediction [e.g., Priebe
The vibro replacement stone-column technique is a popular form of (1995); Castro and Sagaseta (2009), Pulko et al. (2011)]. However,
ground improvement that can be used to enhance the strength and Muir Wood et al. (2000) found that the behavior of small clusters of
stiffness properties of ground intended to support large loaded areas columns supporting footings is far more complex than that of an iso-
(e.g., floor slabs, embankments) or small loaded areas (e.g., foot- lated column. Interaction between the footing and individual columns
ings, strips). Bearing capacity and settlement requirements can be results in a spatial variation in the restraint experienced by columns
satisfied by appropriately choosing the area replacement in plan and beneath the footing; that is, the level of restraint reduces toward the
treatment depth. These factors, and the group size, dictate the mech- footing edge. Muir Wood et al. (2000) conducted a series of labora-
anisms by which columns behave. tory tests to examine the influence of column diameter, length, and
The seminal work by Hughes and Withers (1974) on small-scale spacing on the load-transfer mechanism of small groups of stone col-
physical models identified two modes of deformation for isolated umns. The deformed shape of model stone columns was examined at
stone columns, punching and bulging, with a transition from the for- ultimate conditions using an exhumation technique that uncovered
mer to the latter mechanism as the column length-to-diameter ratio the following four modes of deformation: punching, bulging, bend-
increased beyond 4. The authors also postulated that column bulg- ing, and shearing. Punching was most pronounced in closely spaced
ing can be idealized as a uniform cylindrical expansion, which is a columns, whereas long, widely spaced columns tended to bulge.
The deformational behavior at working stress levels is of
utmost importance because settlement criteria, rather than bear-
1
Lecturer, College of Engineering and Informatics, National Univ. of ing capacity, generally govern foundation design in soft soils
Ireland, Galway H91 HX31, Ireland (corresponding author). E-mail: (Priebe 1976). McKelvey et al. (2004) conducted a series of labora-
bryan.mccabe@nuigalway.ie tory experiments to examine the deformation of columns at opera-
2
Geotechnical Engineer, Arup, Ringsend Road, Dublin D04 T6X0, tional stress levels within a transparent material with clay-like
Ireland; formerly, Research Student, College of Engineering and
properties. Output from three miniature pressure cells was used to
Informatics, National Univ. of Ireland, Galway H91 HX31, Ireland.
Note. This manuscript was submitted on December 2, 2014; approved
ascertain stress-concentration ratios, defined as the quotient of ver-
on March 23, 2016; published online on October 5, 2016. Discussion pe- tical stress in columns and the surrounding soil ðs 0col =s 0soil Þ, and
riod open until March 5, 2017; separate discussions must be submitted for they were correlated with observations of deformed shape. At low
individual papers. This paper is part of the International Journal of applied loads, longer columns bulged and carried a larger propor-
Geomechanics, © ASCE, ISSN 1532-3641. tion of the load ðs 0col =s 0soil  4Þ than shorter columns, which

© ASCE 04016114-1 Int. J. Geomech.

Int. J. Geomech., 04016114


punched into the underlying soil ðs 0col =s 0soil  2Þ. However, (Balaam and Booker 1981), which simplifies the analysis to a single
stress-concentration ratios for both long and short columns tended column and its surrounding zone of influence.
toward a similar value ðs 0col =s 0soil  3Þ at higher applied loads The numerical model was discretized using 15-noded wedge ele-
(i.e., ultimate conditions). This phenomenon is in accordance with ments that were composed of 6-noded triangles and 8-noded quadri-
field measurements, albeit from large groups, by Greenwood laterals. There were three translational degrees of freedom at each
(1991) who reported that stress-concentration ratios decrease with node, and stresses were computed at 6 gauss points located within
applied load when columns are bulging and increase with applied these elements. Symmetry was used to reduce computational time.
load when bulging is precluded. The results of McKelvey et al. The various mesh configurations comprised 14,000–45,000 ele-
(2004) and Greenwood (1991) imply that the mode of deformation ments, which were refined in regions of rapid stress change (i.e.,
at low working loads might differ from that at ultimate conditions. beneath the footing and surrounding the stone columns). The ele-
Black et al. (2011) conducted a series of high-quality physical ments in this region were typically 0.1  0.3 m, which corresponds
model tests to examine the settlement performance of small groups to an aspect ratio of 3. Appropriate mesh-sensitivity studies and
of stone columns. The authors acknowledge the influence of defor- checks on distances to boundaries were also performed to ensure
mational behavior on settlement performance, because they attrib- continuous stress and strain contours at interelement boundaries
Downloaded from ascelibrary.org by Ryerson University on 10/06/16. Copyright ASCE. For personal use only; all rights reserved.

ute the underperformance of a small group of closely spaced (Killeen 2012).


columns, which punch en bloc into the underlying soil, to the load- The deformational behavior of various configurations of stone
transfer mechanism. However, it is difficult to extrapolate the columns was examined by varying the number of columns (N), col-
results of single-gravity model tests because of scale effects and, umn spacing (s), and column length (L). Columns were positioned
therefore, to draw definite conclusions regarding the deformational on a square grid, and the footing width (B) was chosen to extend
behavior at working stress levels. 0.5s beyond the edge of the outer columns. A column diameter (d)
The FEM used in conjunction with advanced constitutive mod- of 600 mm was adopted in the subsequent FEA, because it is
els holds most promise in the mission to understand the link typical for columns formed in soft soils using a standard 430-mm-
between the mode of deformation and settlement performance. diameter poker.
Although previous numerical analyses have been informative, The area replacement ratio, or area ratio, is a parameter that cap-
many are based on plane strain (Muir Wood et al. 2000; Wehr tures the proportion of in situ soil replaced with stone columns. For
2004) or axisymmetric (Elshazly et al. 2008; Hanna et al. 2013) small groups of columns, the area ratio is defined as the ratio of the
conditions, which approximate the presence of stone columns as footing area (A) to the total area of columns beneath the footing (AC)
infinitely long strips or cylindrical rings, respectively. Therefore,
they fail to fully capture the discrete nature of column interaction A 4B2
¼ (1)
observed in the preceding model tests on small groups of columns. AC N p d 2
This shortcoming is overcome by 3D analyses, such as those con-
ducted by Kirsch (2008) and Shahu and Reddy (2011); the Kirsch However, for an infinite grid of columns, as can be assumed for
(2008) study, in particular, showed the effectiveness of this approach large loaded areas modelled using the unit-cell concept, the area ra-
by successfully back-analyzing the primary load-settlement behav- tio is defined as the ratio of the area of the zone of influence of one
ior of a group of five columns in soft alluvial soil. column (A = s2 for a square grid) to the area of a single column (AC)
In a previous paper by the authors (Killeen and McCabe
2014), a parametric study of the factors that affect the settlement A 4s2
¼ (2)
of small groups at working loads in Bothkennar clay, based on a AC p d 2
three-dimensional (3D) finite-element analysis (FEA), was reported.
Drawing on the same output, the influences of key design parameters The area ratio was varied from 3.5 to 14.1, which corresponds to a
such as column length, spacing, and the number of columns on the spacing of 1–2 m to reflect the typical range adopted in practice. The
deformational mechanisms of small groups of stone columns at column length was increased in 1-m increments in the subsequent
working load levels are considered in this paper. The modes of FEA from L = 0 m (i.e., an untreated footing) to L = 13 m to examine
deformation are identified for three specific column lengths through the deformational behavior for partial-depth treatment. The final
an in-depth analysis of the distribution of shear strain within the col- analysis at L = H = 13.9 m (H is the distance from the base of the foot-
umns and soil, in addition to the variation of vertical strain, horizon- ing at a depth of 0.6 m to the base of the Carse clay) represents a full-
tal strain, and stress-concentration ratio along the length of columns. length stone column founded on a stiff (assumed rigid) layer. The
The analysis is extended using novel parameters, referred to as various column group parameters are illustrated for a 3  3 group of
column compression and punching ratios, to determine the influence columns beneath a 3-m-wide square footing in Fig. 1.
of column length and the stiff crust on the mode of deformation.
Hardening Soil Model
Finite-Element Analysis The advanced elastoplastic hardening soil (HS) model is capable of
simulating the behavior of soft and stiff soils (Schanz 1998) and,
hence, was adopted for both the stone columns and surrounding soft
Geometrical Considerations
soil. This model is an extension of the hyperbolic model developed
A 3D FEA was undertaken using PLAXIS 3D Foundation by Duncan and Chang (1970). It supersedes the hyperbolic model
(Brinkgreve and Broere 2006) to investigate the long-term (i.e., because it is based on the theory of plasticity rather than elasticity,
drained) behavior of small groups and infinite grids of stone col- includes soil dilatancy, and introduces a yield cap. The HS model
umns. The small groups of columns were modelled at the center of accounts for the stress-level dependency of soil stiffness by relating
a 3D mesh, the outer boundary of which was laterally restrained but the soil stiffness to the confining stress using a power law, with an
permitted vertical movement. The analysis of an infinite grid of col- exponent m. This model also defines the stiffness response of soil
umns was represented by the conventional unit-cell concept using a secant Young’s modulus at 50% deviatoric stress (E50) and

© ASCE 04016114-2 Int. J. Geomech.

Int. J. Geomech., 04016114


Downloaded from ascelibrary.org by Ryerson University on 10/06/16. Copyright ASCE. For personal use only; all rights reserved.

Fig. 1. Typical layout of a 3  3 group of columns beneath a 3-m-wide square footing in the Bothkennar soil profile

an unload-reload Young’s modulus (Eur). The ultimate strength for It might be argued that the column–soil interlock would not be
the HS model is defined using the Mohr-Coulomb failure criterion, as pronounced for stiff soils and, consequently, a bonded interface
and plastic volumetric strains are captured using the stress-dilatancy would not be appropriate for the stiff crust in this study. However,
theory developed by Rowe (1962). The HS model does not account the low confining stress in this layer due to its proximity to the
for anisotropic strength, the strain-dependent nature of soil stiffness, ground surface enables lateral displacement of the stone backfill
or creep behavior. Further details on the theoretical framework of during column installation. Another important consideration in
this model can be found by Schanz et al. (1999) and Brinkgreve and modeling the column–soil interface is the mode of deformation. For
Broere (2006). example, columns that punch develop larger shear stresses along
the interface, in contrast to columns that bulge. Muir Wood et al.
Other Modeling Issues (2000) cite punching as a dominant mode of deformation for short,
closely spaced columns. Muir Wood et al. (2000) and Wehr (2004,
The load was applied through a 0.6-m-thick footing, which was 2006) successfully recreated the observed mechanisms from these
modelled using a linear elastic material model with typical proper- laboratory tests with the assumption of a bonded interface.
ties of reinforced concrete, i.e., unit weight, g = 24 kN/m3, Young’s Furthermore, Kirsch (2008) and Shahu and Reddy (2011) adopted a
modulus, E = 30 GPa, and Poisson’s ratio, y = 0.15. The footing bonded interface to model load tests on small groups of floating
was loaded to 50 kPa, which is deemed a typical working load on stone columns.
the basis of operational stress levels (20–80 kPa) quoted by
McKelvey et al. (2004).
The long-term behavior of soil can be modelled using either of Development and Validation of Soil Profile
two procedures in PLAXIS 3D Foundation, (1) drained analysis or (2)
undrained loading, followed by consolidation analysis. A 3D FEA on
Bothkennar Soil Profile
groups of 4, 5, and 9 columns (Killeen and McCabe 2014) confirmed
that both procedures produce similar results for stone columns, as did The well-characterized U.K. soft clay test bed at Bothkennar,
an analytical study on an infinite grid of columns by Castro and Scotland, provided the reference soil profile for this research. The
Sagaseta (2011). The drained analysis requires significantly less com- site has been characterized extensively by the Institution of Civil
putational time and was adopted for the subsequent FEA. Engineers (1992) and consists primarily of soft uniform clay
An extensive literature review was undertaken to ascertain the underlying a 1.5-m stiff crust. The Carse clay ranges in thickness
most appropriate modeling procedure for stone columns. In total, from 13 to 20 m across the site and rests on Bothkennar gravel.
24 finite-element studies were reviewed; these studies included sim- Piezometers indicate that the groundwater level is hydrostatic and
ulations of small-scale laboratory tests and field trials for a variety ranges from 0.5 to 1.0 m below ground level (Hight et al. 1992).
of configurations ranging from single columns to large groups. All The Carse clay was deposited in stable marine conditions and, in
of these studies assumed full contact at the column–soil interface, light of findings by Paul et al. (1992) that the clay fraction varies
which is consistent with field observations that stone columns are from 35 to 50%, was classified as silty clay by Nash et al. (1992a).
tightly interlocked with the surrounding soil as a result of the lateral Classification testing by Nash et al. (1992a) also indicates that the
displacement that occurs on relowering the poker during column in- unit weight ( g ) and organic content generally vary from 15.6 to
stallation. McCabe et al. (2009) reported that columns formed using 17.6 kN/m3 and 3 to 8%, respectively. Hight et al. (1992) suggested
the dry bottom-feed process with a standard 430-mm-diameter that postdepositional processes such as erosion, changes in ground-
poker are typically 500–600 mm in diameter. Therefore, to remain water level, and bonding occurred at the Bothkennar test site.
consistent with the approach of several authors (Domingues et al. According to the geological history, these processes could account
2007; Gäb et al. 2008; Elshazly et al. 2008), a bonded interface was for a 15-kPa drop in effective overburden pressure, which explains
adopted along the column–soil interface. the high coefficients of lateral earth pressure at rest in the upper

© ASCE 04016114-3 Int. J. Geomech.

Int. J. Geomech., 04016114


layers observed by Nash et al. (1992a). However, the lower layers The HS model material parameters adopted for the Bothkennar
might be influenced by aging, because the in situ stress state is better soil profile are summarized in Table 1. The value of E50 for the stiff
represented by an overconsolidation ratio of 1.5. crust quoted in this table seems to be quite low; however, it should
The mechanical properties of reconstituted Carse clay were be noted that this value is specific to pref = 13 kPa. Normalizing E50
examined by Allman and Atkinson (1992). Soil samples were for pref = 100 kPa (the default setting in PLAXIS 3D Foundation)
reconstituted, formed into slurry (at 1.25 times the liquid limit), and yields E50 = 8,215 kPa. Furthermore, the stiff crust is influenced sig-
loaded to normally consolidated, or swelled to lightly overconsoli- nificantly by Eur (which is equal to 41,398 kPa for pref = 100 kPa)
dated, stress states. The samples were then transferred to the triaxial for the initial stages of loading because of its overconsolidated
cell for testing, and a critical-state friction angle of f 0 = 34° in com- stress state.
pression was measured. This angle is larger than would be expected
for high-plasticity clay and might be a result of the high proportion
of angular silt particles. No effective cohesion was observed in the Simulation of Load Test on Bothkennar Soil Profile
tests; however, nominal values were adopted to ensure numerical A field trial on small groups of stone columns was conducted at
stability in the FEA. the Bothkennar test site by Serridge and Sarsby (2008).
Downloaded from ascelibrary.org by Ryerson University on 10/06/16. Copyright ASCE. For personal use only; all rights reserved.

Nash et al. (1992b) undertook a comprehensive study to examine However, the authors reported that the tests contained signifi-
the 1D stiffness properties of Carse clay. The variation of the coeffi- cant installation effects because excessive air jetting and com-
cient of compressibility (CC) with depth is amply documented, and paction pressure were applied to form an enlarged bulb at the
the Bothkennar soil profile can be subdivided into three strata on the base of stone columns. Therefore, it was not appropriate to vali-
basis of these data, namely, crust, upper Carse clay, and lower date the finite-element model using these field data. Instead, the
Carse clay. However, the behavior before yielding was not meas- adopted soil profile and material parameters were validated by
ured routinely, and reference was made to Allman and Atkinson simulating a field test on an unreinforced footing at the
(1992), who reported the ratio between the slopes of the normal Bothkennar test site using PLAXIS 3D Foundation. Field load
compression (l ) and swelling lines (k ) to be 7.2, which thereby tests were conducted on two rigid pad footings documented by
allows the coefficient of swelling (CS) to be determined, i.e., l /k = Jardine et al. (1995). In Fig. 2, a simulation of the short-term
CC/CS = 7.2. The 1D stiffness properties, adapted from incremental
load test (i.e., without significant creep settlement) is com-
load-test data, may then be converted into Eoed and Eur for the HS
pared with the observed load-displacement curve. Because of
model, using the following expressions outlined by Brinkgreve and
the short-term nature of the loading, the Carse clay is mod-
Broere (2006):
elled as an undrained material using an effective stress analy-
2:3ð1 þ e0 Þpref sis with pore water pressure. Effective strength properties
ref
Eoed ¼ (3) were used in conjunction with the HS model for this analysis.
CC
This procedure replicates undrained conditions by adding a
large bulk modulus for the pore water pressure to the stiffness
2:3ð1 þ e0 Þð1 þ  Þð1  2 Þpref matrix to generate excess pore pressure. As an alternative, the
ref
Eur ¼ (4)
CS ð1   Þ undrained shear strength could have been defined directly, but
it would have required the stress dependency of soil stiffness
where Eref = reference stiffness corresponding to the reference con- and compression-hardening features of the HS model to be
fining pressure (pref). deactivated. It can be seen in Fig. 2 that PLAXIS 3D Foundation
The stress-level dependency of the Carse clay stiffness can be captures the variation of settlement with foundation pressure
captured accurately using a power law with an exponent m = 1.0 quite well, although it overestimates settlement slightly at low
(Killeen 2012), which is typical of soft clay. Although no direct applied pressures. This may be attributed in part to the presence
relationship exists between Eoed and E50, it was deemed reasonable of a localized bed of shelly fragments at the base of the crust,
to assume a relationship as outlined in Eq. 5 (Brinkgreve and which would have enhanced the stiffness response of the field
Broere 2006) test without affecting the bearing capacity. Overall, this result
gives confidence in the combination of soil model and material
ref
E50 ¼ 1:25Eoed
ref
(5) parameters.

Table 1. Summary of Material Parameters Adopted for Bothkennar Test Site and Stone Column

Soil parameter Crust Upper Carse clay Lower Carse clay Stone column
Depth (m) 0.0–1.5 1.5–2.5 2.5–14.5 —
Bulk unit weight [ g (kN/m3)] 18.0 16.5 16.5 19.0
Overconsolidation ratio — — 1.5 —
Pre-overburden stress (kPa) 15 15 — —
Coefficient of lateral earth pressure (K0) 1.5 1.0 0.75 0.29
Secant Young’s modulus at 50% deviatoric stress [E50 (kPa)] 1,068 506 231 70,000
Unload-reload Young’s modulus [Eur (kPa)] 5,382 3,306 1,164 210,000
Reference pressure [pref (kPa)] 13 20 30 100
Stress dependency power-law exponent (m) 1.0 1.0 1.0 0.3
Poisson’s ratio (y ) 0.2 0.2 0.2 0.2
Effective cohesion ½c0 ðkPaÞ 3 1 1 1
Angle of internal friction ½ f 0 ð Þ 34 34 34 45
Angle of dilation ½ c 0 ð Þ 0 0 0 15

© ASCE 04016114-4 Int. J. Geomech.

Int. J. Geomech., 04016114


140

120
Jardine et al. (1995) PLAXIS 3D Foundation

Applied pressure (kPa)


100

80

60

40

20
Downloaded from ascelibrary.org by Ryerson University on 10/06/16. Copyright ASCE. For personal use only; all rights reserved.

0
0 20 40 60 80 100 120 140 160 180 200
Vertical displacement (mm)

Fig. 2. Comparison of PLAXIS 3D Foundation with actual load-displacement behavior for a pad footing at Bothkennar test site observed by Jardine
et al. (1995)

Stone Backfill respectively. The modes of deformation inferred from these plots
are discussed later.
The bulk unit weight of stone columns was chosen as g = 18.6 kN/m3,
which is consistent with that of previous studies (Mitchell and Huber Column Punching
1985; Domingues et al. 2007; and Gäb et al. 2008). On the basis of a
It can be seen in Fig. 3 that the majority of shear strain develops
settlement database of field load tests on stone columns compiled
beneath the base of 3-m-long columns, which indicates that punch-
by McCabe et al. (2009) and a laboratory study by Herle et al. (2006),
ing is the dominant mode of deformation. Moreover, the distribu-
f 0 = 45° was adopted for stone columns. The angle of dilation ( c )
tion of shear strain beneath the base of the group is more uniform at
was determined from the empirical relationship c 0 = f 0 – 30°, as
low area ratios and becomes more localized around the column
suggested by Brinkgreve and Broere (2006). The stiffness param-
bases with increasing area ratios, which suggests that short, closely
eters are similar to those adopted by Gäb et al. (2008) (i.e., E50 =
spaced columns tend to punch as a block [Fig. 3(a)], whereas short,
70 MPa, Eur = 210 MPa, and m = 0.3), who modelled stone
widely spaced columns tend to punch individually [Fig. 3(c)]. The
columns in loose–medium dense sand and weak clayey silt beneath
high levels of shear strain generated along the outer face and
an embankment. These values are higher than Young’s moduli (E =
beneath the base of floating columns in Figs. 3(a) and 4(a) shows
30–58 MPa) reported by Mitchell and Huber (1985) but are justified
that punching, in the form of a block, remains the dominant mode of
by the fact that columns formed by the newer bottom-feed process
deformation for all floating columns at low area ratios. Block
were found by McCabe et al. (2009) to yield an enhanced settlement
punching is a form of column punching and is characterized by low
performance.
levels of shear strain in the central zone of soil bounded by columns.
This mode of deformation was also observed by Black et al. (2011)
Results of Numerical Analysis for a closely spaced group of three columns (A/AC = 2.5 and 3.6) in
small-scale laboratory tests.
The mode of deformation for different configurations of columns was
identified by examining the distribution of shear strain within columns Column Bulging
and the surrounding soil. In addition, the characteristic column behav- It can be seen in Figs. 4 and 5 that the mode of deformation transits
ior was also examined through plots of horizontal strain, vertical from punching to bulging with increasing area ratios in the case of
strain, and stress-concentration ratio along the length of the columns. 8- and 13.9-m-long columns. Bulging is most pronounced at the top
Using this approach, Killeen (2012) analyzed the deformational of the lower Carse clay layer, because this level corresponds to the
behavior for the following commonly used group configurations for weakest part of the soil profile. Bulging does not occur with increas-
square footings: single columns, 2  2, 3  3, and 4  4 groups, and ing area ratios for 3-m-long columns because of the high levels of
infinite grids. The various modes of deformation identified were all in restraint provided by the stiff upper layers.
evidence for a 3  3 group at different lengths and spacings.
Therefore, for the purposes of brevity, the results presented herein are Transition from Punching to Bulging
confined to those of a 3  3 group of columns at lengths of 3, 8, and The transition from punching to bulging with increasing area ratios
13.9 m. The analysis was extended using newly defined compression for 8- and 13.9-m-long columns might be explained by the increase
and punching ratios to investigate the deformational behavior of all in load carried per column at higher area ratios. Hughes and
configurations and, also, the influence of the stiff crust. Withers (1974) showed that the occurrence of punching or bulging
was contingent on the relative contributions of side friction, base re-
sistance, and lateral resistance experienced by stone columns. In
Distribution of Total Shear Strain
short, bulging tends to occur if the combined side friction and base
The distributions of shear strain for a 3  3 group of columns at resistance exceed the lateral resistance provided to columns; other-
lengths of 3, 8, and 13.9 m are shown in Figs. 3, 4, and 5, wise, punching occurs. It is well known from piling theory that side

© ASCE 04016114-5 Int. J. Geomech.

Int. J. Geomech., 04016114


(×10-3) (×10-3) (×10-3)
Crust

Upper Carse clay

Lower Carse clay

Uniform strain at
base of columns

Localised punching
at column bases
Downloaded from ascelibrary.org by Ryerson University on 10/06/16. Copyright ASCE. For personal use only; all rights reserved.

(a) (b) (c)

Fig. 3. Total shear strains for a 3  3 group of 3-m-long columns at area ratios (A/AC) of (a) 3.5, (b) 8.0, and (c) 14.1

(×10-3) (×10-3) (×10-3)


Crust

Upper Carse clay

Lower Carse clay

Low γ in soil
bounded by
outer columns

Bulging occurs at top of


Lower Carse clay

(a) (b) (c)

Fig. 4. Total shear strains for a 3  3 group of 8-m-long columns at area ratios (A/AC) of (a) 3.5, (b) 8.0, and (c) 14.1

friction and base resistance become fully mobilized at low and high corner of the pad footing and the point of maximum bulging, which
strain levels, respectively. Therefore, as the magnitude of applied is akin to a general shear failure. A similar mechanism was
pressure increases, the combined effects of side friction and base re- observed in small-scale model tests by Muir Wood et al. (2000) and
sistance also increase. As a result, punching is most likely at low confirmed in a FEA by Wehr (2004). However, the shear planes in
levels of applied pressure when the combined effects of side friction these studies were more pronounced and intersected to bound a
and base resistance are low, whereas bulging is most likely at high cone of undeforming soil beneath the base of footings. This result
levels of applied pressure. This is consistent with the work of may be explained by the fact that these studies correspond to ulti-
McKelvey et al. (2004) who observed that the vertical stress carried mate conditions, at which point general shear failure tends to occur.
by punching columns increases with the level of applied pressure The authors observed that the footing width was one of the principal
before ultimately reaching a similar value to that for bulging col- factors affecting the extent of this zone and tentatively suggested a
umns, which implies that the mode of deformation of columns critical length of 1.5B, whereby columns shorter and longer than
depends not only on the geotechnical capacity of columns but also this length exhibit punching and bulging, respectively. This finding
on the magnitude of applied pressure. seems to be specific to ultimate conditions, because it is not sup-
ported in current numerical analysis at working load levels; in fact,
Secondary Mode of Deformation the opposite trend can be seen in Figs. 4 and 5 as the mode of defor-
It can also be seen in Figs. 4 and 5 that a secondary mode of defor- mation shifts from punching to bulging with a reduction in L/B
mation begins to develop in the form of a shear plane between the [e.g., Figs. 4(a–c)]. This result adds further evidence to suggest that

© ASCE 04016114-6 Int. J. Geomech.

Int. J. Geomech., 04016114


(×10-3) (×10-3) (×10-3)
Crust Shearing
Upper Carse clay

Lower Carse clay

Bulging is primary
mode of deformation
Downloaded from ascelibrary.org by Ryerson University on 10/06/16. Copyright ASCE. For personal use only; all rights reserved.

(a) (b) (c)

Fig. 5. Total shear strains for a 3  3 group of 13.9-m-long end-bearing columns at area ratios (A/AC) of (a) 3.5, (b) 8.0, and (c) 14.1

the parameters that govern the mode of deformation vary with the Column Bulging
magnitude of applied pressure. Column bulging is readily identified for 8-m-long columns in
Figs. 7(a and b) and 13.9-m-long columns in Figs. 8(a and b) at
A/AC values of 8.0 and 14.1 by high levels of vertical and horizon-
Characteristic Column Behavior tal strain within columns, respectively. The magnitude and extent
The characteristic column behavior for a 3  3 group is examined of strain both increase with the area ratio and column length,
through plots of vertical strain (« y), horizontal strain (« h), and the which reflects the increased load carried per column and
stress-concentration ratio ðs 0col =s 0soil Þ along the length of 3-, 8-, and enhanced load-carrying capacity of columns, respectively. The
13.9-m-long columns. PLAXIS 3D Foundation determines strain location of maximum strain, however, remains constant at a depth
along three orthogonal axes (i.e., x, y, and z); « y = strain along the of approximately 2.5 m, which is close to the top of the lower
vertical axis, and « h = average of the two orthogonal strains along Carse clay (i.e., the point of lowest lateral restraint). Stress-
the horizontal axis [i.e., « h = (1/2)(« x þ « z)]. The stress-concentra- concentration ratios are closely correlated to the distribution of
tion ratio ðs 0col =s 0soil Þ presented in the following section compares strain, because s 0col =s 0soil decreases from the ground surface to a
the vertical effective stress in columns to the average vertical effec- local minimum at the point of maximum bulging. Sections of
tive stress in the soil beneath the footprint area of the footing. The columns that are bulging are in a contained state of plasticity,
stresses (s 0 ) in this quotient refer to the cumulative total of the in which reduces their load-carrying capacity. Stress-concentration
situ stress ðs 00 Þ and the increment of stress that results from applied ratios increase temporarily below this point because of the
loading ðDs 0 Þ (i.e., s 0 ¼ s 00 þ Ds 0 ); initially, before the applica- enhanced levels of confinement associated with increasing over-
tion of footing pressure, s 0col =s 0soil is close to unity. The results pre- burden stress, before decreasing with depth in sections of col-
sented herein are the average of the corner, edge, and center umns that do not yield, which again reflects the increasing influ-
columns. ence of the in situ stress component in s 0col =s 0soil .

Column Punching
Compression and Punching Ratios
It was established in the previous section that punching was the
dominant mode of deformation for short columns and closely Definition of Compression and Punching Ratios
spaced columns. These configurations are characterized by high As illustrated in Fig. 9, the compression and punching ratios are
vertical strain beneath the base of floating columns [Figs. 6(a) and 7 derived from displacement at the surface of footings (ufooting), the
(a)] and negligible horizontal strain within columns [Figs. 6(b) and base of columns (ucol), and the soil surrounding the base of columns
7(b)]. It is interesting to note that 3-m-long columns at higher area (usoil) averaged over a 1-m-square zone. The width of this zone also
ratios attempt to bulge, as evidenced by the increase in horizontal could have corresponded to the column spacing, but it was felt that
strain shown in Fig. 6(b), but punching remains the primary mode selecting a constant width ensured a consistent means of compari-
of deformation. It can be seen in Figs. 6(c), 7(c), and 8(c) that son for all column configurations (Killeen 2012).
stress-concentration ratios for short columns and closely spaced The compression ratio is defined as the column compression
columns are highest at the ground surface and decrease steadily to- (ufooting – ucol) normalized by ufooting [Eq. (6)]. This ratio is essen-
ward unity at the base of columns. This trend is a result of the reduc- tially a measure of the axial strain along the length of columns. It
tion of vertical stress in columns through side friction and, also, as varies from 0 to 1 in the case of perfectly rigid columns to fully
the in situ stress component in s 0col =s 0soil becomes larger with compressible or end-bearing columns, respectively. Muir Wood
increasing depth. These stress-concentration ratios are at the upper et al. (2000) and Black et al. (2011) measured the load transfer to
end of a typical range (2.5–5.0) quoted by Barksdale and Bachus the base of columns using a related ratio, ucol/ufooting, which allows
(1983). for a comparison of results.

© ASCE 04016114-7 Int. J. Geomech.

Int. J. Geomech., 04016114


Vertical strain, εy (× 10-3) Horizontal strain, εh (× 10-3) Stress concentration, σ'col/σ'soil
0 -10 -20 -30 -40 0 5 10 15 20 1 3 5 7 9
0

4
Depth, z (m)

8
3×3 cols 3×3 cols 3×3 cols
10 (average) (average) (average)
A/Ac = 3.5 A/Ac = 3.5 A/Ac = 3.5
Downloaded from ascelibrary.org by Ryerson University on 10/06/16. Copyright ASCE. For personal use only; all rights reserved.

12 A/Ac = 8.0 A/Ac = 8.0 A/Ac = 8.0


A/Ac = 14.1 A/Ac = 14.1 A/Ac = 14.1
14

(a) (b) (c)

Fig. 6. Distributions of the average (a) vertical strain, (b) horizontal strain, and (c) stress-concentration ratio with depth for a 3  3 group of 3-m-long
columns

Vertical strain, εy (× 10-3) Horizontal strain, εh (× 10-3) Stress concentration, σ'col/σ'soil


0 -10 -20 -30 -40 0 5 10 15 20 1 3 5 7 9
0

2
σ'col/σ'soil ↓ in
4 section which
is bulging
Depth, z (m)

8
3×3 cols 3×3 cols 3×3 cols
10 (average) (average) (average)
A/Ac = 3.5 A/Ac = 3.5 A/Ac = 3.5
12 A/Ac = 8.0 A/Ac = 8.0 A/Ac = 8.0
A/Ac = 14.1 A/Ac = 14.1 A/Ac = 14.1
14
(a) (b) (c)

Fig. 7. Distributions of the average (a) vertical strain, (b) horizontal strain, and (c) stress-concentration ratio with depth for a 3  3 group of 8-m-long
columns

ufooting  ucol These ratios should be considered in tandem to assess the mode
compression ratio ¼ (6)
ufooting of deformation. For example, a punching column will yield low
compression and high punching ratios because a large proportion of
A limitation of the compression ratio is that it does not capture the load is transferred to the base, whereas a bulging column will
column punching properly, which is cited as the prevalent mode yield high compression and low punching ratios because load trans-
of deformation for short columns and closely spaced columns in fer to the base is limited by excessive radial deformation of the
various model studies (Muir Wood et al. 2000; McKelvey et al. column.
2004; Black et al. 2011). Therefore, it is necessary to introduce a
punching ratio that quantifies the proportion of applied load trans- Modes of Deformation
ferred to the base of columns. This ratio is defined as the relative The modes of deformation are discussed in terms of the influences
displacement at the base of columns (ucol – usoil), again normal- of column length and group size with respect to area ratio. The influ-
ized by ufooting [Eq. (7)] ence of column length is split into two categories: columns with an L
value of ≤3 m are influenced primarily by the stiff upper layers,
ucol  usoil whereas columns with an L value of >3 m are influenced primarily
punching ratio ¼ (7)
ufooting by the soft clay. The key observations are summarized next.

© ASCE 04016114-8 Int. J. Geomech.

Int. J. Geomech., 04016114


Vertical strain, εy (× 10-3) Horizontal strain, εh (× 10-3) Stress concentration, σ'col/σ'soil
0 -10 -20 -30 -40 0 5 10 15 20 1 3 5 7 9
0

2
σ'col/σ'soil ↓ in
4 section which
is bulging
Depth, z (m)

8
3×3 cols 3×3 cols 3×3 cols
10 (average) (average) (average)
A/Ac = 3.5
Downloaded from ascelibrary.org by Ryerson University on 10/06/16. Copyright ASCE. For personal use only; all rights reserved.

A/Ac = 3.5 A/Ac = 3.5


12 A/Ac = 8.0 A/Ac = 8.0 A/Ac = 8.0
A/Ac = 14.1 A/Ac = 14.1 A/Ac = 14.1
14
(a) (b) (c)

Fig. 8. Distributions of the average (a) vertical strain, (b) horizontal strain, and (c) stress-concentration ratio with depth for a 3  3 group of 13.9-m-long
end-bearing columns

Fig. 9. Displacement at surface of footing, base of columns, and soil surrounding base of columns

Column Length, L ≤ 3 m. Column Length, L > 3 m.


• Punching is the dominant mode of deformation. It can be seen • Punching remains the dominant mode of deformation at low
in Figs. 10(a), 11(a), and 12(a) that punching ratios increase A/AC values. With the exception of an infinite grid of columns
with column length up to a maximum at L = 3 m. These ratios are at an A/AC value of 3.5, it can be seen in Fig. 10(a) that punch-
coupled with low compression ratios, as shown in Figs. 10(b), ing ratios reduce with increasing column length, which is a
11(b), and 12(b), which suggests that columns do not deform result of an enhanced side friction and stiffer soil at the base of
appreciably along their length, which confirms a punching mode columns. The reduction in punching ratios is coupled with an
of deformation. increase in compression ratios, as shown in Fig. 10(b). The
• Punching ratios increase with column length. The increase in rate of change of these ratios with column length is important
punching ratios with column lengths up to L = 3 m can be in determining the mode of deformation. For example, it can
attributed to a progression through the layers of reducing stiff- be seen in Figs. 11(b) and 12(b) that the increase in compres-
ness (i.e., the crust is the stiffest layer) followed by the upper sion ratios with column length occurs at a faster rate for higher
Carse clay and then the lower Carse clay (as reflected in the CC area ratios. This reflects a shift in the mode of deformation
and K0 values shown in Table 1). Higher punching ratios occur from punching to bulging for widely spaced columns longer
when the base of columns are founded in more compressible than 3 m, and the latter mechanism is characterized by exces-
strata, as shown in the PLAXIS 3D Foundation output for an sive radial deformation. In contrast, the slower increase in
infinite grid of columns in Fig. 13. It is interesting to note that compression ratios with column length for closely spaced col-
maximum punching ratios occur for L = 3 m, despite the fact umns shown in Fig. 10(b) suggests that the mode of deforma-
that the base of 2-m-long columns coincides with the weakest tion observed for short columns (i.e., punching) continues to
part of the soil profile. This may be explained as the stressed occur with increasing column length. This result is consistent
zone of soil surrounding the base of 2-m-long columns extends with a punching mode of deformation and agrees with the dis-
upward into the overlying (stiffer) upper Carse clay, which tribution of shear and normal strains presented in the previous
thereby reduces punching ratios. sections. This indicates that punching remains the dominant

© ASCE 04016114-9 Int. J. Geomech.

Int. J. Geomech., 04016114


Punching ratio, (ucol-usoil)/ufooting Compression ratio, (ufooting-ucol)/ufooting
0.00 0.05 0.10 0.15 0.20 0.25 0.0 0.2 0.4 0.6 0.8 1.0
0 0
A/AC = 3.5
2 2 1 col
2×2
Column length, L (m)

Column length, L (m)


3×3 - average
4 4 4×4 - average
Unit Cell
6 6

8 8
A/AC = 3.5
10 1 col 10
Downloaded from ascelibrary.org by Ryerson University on 10/06/16. Copyright ASCE. For personal use only; all rights reserved.

2×2
12 3×3 - average 12
4×4 - average
Unit Cell
14 14

(a) (b)

Fig. 10. Influence of column confinement on (a) punching and (b) compression ratios for various lengths and configurations of columns spaced at an
area ratio (A/AC) of 3.5

Punching ratio, (ucol-usoil)/ufooting Compression ratio, (ufooting-ucol)/ufooting


0.00 0.05 0.10 0.15 0.20 0.25 0.0 0.2 0.4 0.6 0.8 1.0
0 0
A/AC = 8.0
2 2 1 col
2×2
Column length, L (m)

Column length, L (m)

3×3 - average
4 4 4×4 - average
Unit Cell
6 6

8 8
A/AC = 8.0
10 1 col 10
2×2
12 3×3 - average 12
4×4 - average
Unit Cell
14 14

(a) (b)

Fig. 11. Influence of column confinement on (a) punching and (b) compression ratios for various lengths and configurations of columns spaced at an
area ratio (A/AC) of 8.0

mode of deformation with increasing column lengths for a for L = 3 m, as illustrated by plots of the deformed shape for an
group of closely spaced columns. infinite grid of columns in Fig. 14. It can be seen in Fig. 14(a)
• Bulging mode of deformation occurs at high A/AC values. The that columns at a low area ratio act with the surrounding soil
decrease in punching ratios and increase in compression ratios and punch as a single entity into the underlying soil. This is
are most pronounced for higher area ratios (Figs. 11 and 12). In referred to as block punching, which is frequently observed for
effect, widely spaced columns transfer a smaller proportion of closely spaced groups of piles. In contrast, column interactions
the applied load to their base and exhibit increased axial defor- diminish at a higher area ratio because columns tend to act in
mation, which again is consistent with plots of shear and normal increasing isolation from each other [Figs. 14(b and c)], which
strain outlined in the previous sections. yields higher punching ratios, as observed in Figs. 11(a) and
Influence of the Group Size. 12(a).
• Block punching occurs for larger groups at low A/AC values.
It is evident in Fig. 10(a) that the magnitude of punching ratio Implications for Design of Stone Columns
is reduced significantly for larger groups at low area ratios. A simplified design method, which relates the settlement of small
This is evident for all column lengths and is most pronounced groups (d ) to that of a unit cell (d uc) with knowledge of the area

© ASCE 04016114-10 Int. J. Geomech.

Int. J. Geomech., 04016114


Punching ratio, (ucol-usoil)/ufooting Compression ratio, (ufooting-ucol)/ufooting
0.00 0.05 0.10 0.15 0.20 0.25 0.0 0.2 0.4 0.6 0.8 1.0
0 0
A/AC = 14.1
2 2 1 col
2×2
Column length, L (m)

Column length, L (m)


3×3 - average
4 4 4×4 - average
Unit Cell
6 6

8 8
A/AC = 14.1
10 1 col 10
Downloaded from ascelibrary.org by Ryerson University on 10/06/16. Copyright ASCE. For personal use only; all rights reserved.

2×2
12 3×3 - average 12
4×4 - average
Unit Cell
14 14

(a) (b)

Fig. 12. Influence of column confinement on (a) punching and (b) compression ratios for various lengths and configurations of columns spaced at an
area ratio (A/AC) of 14.1

Fig. 13. Deformation of an infinite grid of floating stone columns at an area ratio (A/AC) of 14.1 for various column lengths: (a) L = 1 m; (b) L = 2 m;
(c) L = 3 m

ratio, column length, footing width, and depth of soft soil, was Influence of Stiff Upper Layers
proposed by Killeen and McCabe (2014). This study has shown The laboratory studies referenced in the previous sections were con-
the importance of area ratio and column length in dictating the ducted in homogenous test beds that did not include a stiff crust.
mechanism by which groups behave. It is important to note that This is salient feature of many soft-soil profiles, and its importance
the study also shows that the reason why a simple d /d uc ratio at the Bothkennar test site was highlighted by Serridge and Sarsby
approach can be adopted is that the unit cell and groups show fun- (2008), who conducted load tests on stone column–reinforced pad
damentally similar mechanisms of behavior (Figs. 10, 11, and footings. Therefore, to put the findings from previous laboratory
12). Therefore, group settlements that are derived from d uc values studies into context, an additional analysis was conducted to investi-
inherently capture the appropriate mechanism. The only differ- gate the effect of the stiff upper layers. The deformational behavior
ence between the unit-cell and group configurations arises from of a 2  2 group of columns at an A/AC value of 8.0 was examined
the stress distribution beneath the footing, which allows the well- for a uniform soil profile, which consisted entirely of lower Carse
established design methods developed for unit-cell conditions clay (i.e., with the crust removed).
[e.g., Priebe (1995); Castro and Sagaseta (2009); and Pulko et al. The variations of punching and compression ratios with column
(2011)] to be used in conjunction with an empirical correction length are presented in Figs. 15(a and b), respectively, along with
based on geometry, to determine the settlement of a stone-column data adapted from Muir Wood et al. (2000) and Black et al. (2011).
group. The uniform soil profile yields high punching ratios for short

© ASCE 04016114-11 Int. J. Geomech.

Int. J. Geomech., 04016114


Downloaded from ascelibrary.org by Ryerson University on 10/06/16. Copyright ASCE. For personal use only; all rights reserved.

Fig. 14. Deformation of an infinite grid of 3-m-long columns at various area ratios (A/AC): (a) 3.5; (b) 8.0; (c) 14.1

Punching ratio, (ucol - usoil)/usurface Compression ratio, (usurface - ucol)/usurface


0.0 0.1 0.2 0.3 0.4 0.0 0.2 0.4 0.6 0.8 1.0
0 Normalised column length, L/d 0
Normalised column length, L/d

Absence of crust
5 increases column 5
punching for short
columns
10 10
Absence of crust reduces load
transfer to base of long columns Absence of crust increases
column compression
15 15

20 20 Black et al. (2011)


Muir Wood et al. (2000)
PLAXIS: 4 cols; A/AC=8.0 PLAXIS: 4 cols; A/AC=8.0
PLAXIS: 4 cols; A/AC=8.0; no crust PLAXIS: 4 cols; A/AC=8.0; no crust
25 25
(a) (b)

Fig. 15. Influence of stiff upper layers on (a) punching and (b) compression ratios for various lengths of a 2  2 group of columns at an A/AC = 8.0

columns that decay sharply with depth. This is coupled with rela- stone column–reinforced foundations, because longer columns
tively high compression ratios, which is indicative of a bulging (previously assumed to be ineffective for settlement control) might
mode of deformation. Compression ratios for the uniform soil pro- be used as a viable alternative configuration.
file agree quite well with laboratory data for small groups at A/AC
values of 3.3–10 (Muir Wood et al. 2000) and single columns at
A/AC values of 2.5–5.9 (Black et al. 2011). However, the model Conclusions
tests yield compression ratios of unity for L/d > 10, which are
slightly higher than those from PLAXIS 3D Foundation and might The deformational behavior of various configurations of columns
be attributed to increased bulging (and reduced load transfer to the beneath rigid square pad footings in the Bothkennar soil profile has
base of columns) at ultimate conditions. This has prompted authors been examined in this study and found to be a complex process. The
in the past to suggest the existence of a unique critical length. This conclusions are summarized in three categories, as follows.
finding is not supported in the current analysis for the reference soil
profile at working load levels. Instead, it appears that a combination
Mode of Deformation
of punching and bulging contributes to the deformed shape of col-
umns, and bulging becomes more prominent at higher loads. This is The following modes of deformation were identified: individual
an important finding that gives greater flexibility when designing punching, block punching, and bulging.

© ASCE 04016114-12 Int. J. Geomech.

Int. J. Geomech., 04016114


• Individual and block punching are related, because high strain Domingues, T. S., Borges, J. L., and Cardoso, A. S. (2007). “Parametric
levels develop along the outer face and beneath the base of col- study of stone columns in embankments on soft soils by finite element
umns. These modes of deformation typically develop in short method.” Applications of Computational Mechanics in Geotechnical
or closely spaced columns, and block punching is more preva- Engineering V: Proc., 5th Int. Workshop, CRC Press, Boca Raton, FL,
281–291.
lent in larger groups.
Duncan, J. M., and Chang, C. Y. (1970). “Nonlinear analysis of stress and
• Bulging occurs for long, widely spaced columns and highly strain in soil.” J. Soil Mech. Found. Div., 96(5), 1629–1653.
depends on the lateral resistance provided by the surrounding Elshazly, H. A., Hafez, D. H., and Mossaad, M. E. (2008). “Reliability of
soil. Columns bulge at shallow depths (i.e., the point of lowest conventional settlement evaluation for circular foundations on stone col-
lateral resistance), which limits load transfer to the base of umns.” J. Geotech. Geol. Eng., 26(3), 323–334.
columns. Gäb, M., Schweiger, H. F., Kamrat-Pietraszewska, D., and Karstunen, M.
• At working stress levels, the occurrence of punching or bulg- (2008). “Numerical analysis of a floating stone column foundation using
ing is dictated by the area ratio and column length rather than different constitutive models.” Proc., 2nd Int. Workshop on Geotechnics
the number of columns. of Soft Soils, 137–142.
Greenwood, D. A. (1991). “Load tests on stone columns.” STP 1089: Deep
foundation improvements: Design, construction and testing, ASTM,
Downloaded from ascelibrary.org by Ryerson University on 10/06/16. Copyright ASCE. For personal use only; all rights reserved.

Stress-Concentration Ratios West Conshohocken, PA, 141–171.


• For punching columns, stress-concentration ratios are high- Hanna, A. M., Etezad, M., and Ayadat, T. (2013). “Mode of failure of a
group of stone columns in soft soil.” Int. J. Geomech., 10.1061
est at the ground surface and decrease steadily toward unity
/(ASCE)GM.1943-5622.0000175, 87–96.
at the base of columns, which reflects the reduction in verti- Herle, I., Hentschel, M., Wehr, W., and Bohac J. (2006). “Role of density
cal stress through side friction and increasing in situ stress on the behaviour of vibrated stone columns in soft soils.” Proc., Int.
with depth. Symp. on Vibratory Pile Driving & Deep Soil Vibratory Compaction,
• For bulging columns, stress-concentration ratios are closely Laboratoire Central des Ponts et Chaussées, Paris, 141–146.
related to the distribution of strain within columns, because Hight, D. W., Bond, A. J., and Legge, J. D. (1992). “Characterization of the
they decrease from the surface to a local minimum at the Bothkennaar clay: An overview.” Geotechnique, 42(2), 303–347.
point of maximum bulging. Stress-concentration ratios Hughes, J. M., and Withers, N. J. (1974). “Reinforcing of soft cohesive soils
increase temporarily with depth, because of increasing con- with stone columns.” Ground Eng., 7(3), 42–49.
Institution of Civil Engineers. (1992). “Bothkennar soft clay test site:
finement with overburden stress, before tending toward
Characterization and lessons learned.” Geotechnique, 42(2), 161–378.
unity. Jardine, R. J., Lehane, B. M., Smith, P. R., and Gildea, P. A. (1995).
“Vertical loading experiments on rigid pad foundations at Bothkennar.”
Influence of Stiff Upper Layers Geotechnique, 45(4), 573–597.
Killeen, M. M. (2012). “Numerical modelling of small groups of stone col-
• The absence of a stiff crust increases the likelihood of column umns.” Ph.D. thesis, National Univ. of Ireland, Galway, Ireland.
bulging at shallower depths and, consequently, reduces the Killeen, M. M., and McCabe, B. A. (2014). “Settlement performance of pad
load transfer to the base of columns. This might have prompted footings on soft clay supported by stone columns: A numerical study.”
authors in the past to suggest the existence of a unique critical Soils Found., 54(4), 760–776.
length on the basis of laboratory studies conducted in homoge- Kirsch, F. (2008). “Evaluation of ground improvement by groups of
neous test beds. However, the presence of a stiff crust forces vibro stone columns using field measurements and numerical analy-
bulging to occur deeper and increases the load transfer to the sis.” Proc., 2nd Int. Workshop on the Geotechnics of Soft Soils,
base of columns, which enables greater flexibility when CRC Press, Boca Raton, FL, 241–248.
McCabe, B. A., Nimmons, G. J., and Egan, D. (2009). “A review of field
designing stone column–reinforced foundations.
performance of stone columns in soft soils.” Proc. Inst. Civ. Eng.
In addition, the general compatibility between the mechanisms of Geotech. Eng., 162(6), 323–334.
behavior for unit cells and groups has justified the use of design McKelvey, D., Sivakumar, V., Bell, A. L., and Graham, J. (2004).
methods relating the settlement of groups to that of a unit cell, such “Modelling vibrated stone columns in soft clay.” Proc. Inst. Civ. Eng.
as that reported by Killeen and McCabe (2014). Geotech. Eng., 157(3), 137–149.
Mitchell, J. K., and Huber, T. R. (1985). “Performance of a stone column
foundation.” J. Geotech. Eng., 10.1061/(ASCE)0733-9410(1985)111:
References 2(205), 205–223.
Muir Wood, D., Hu, W., and Nash, D. F. T. (2000). “Group effects in stone
Allman, M. A., and Atkinson, J. H. (1992). “Mechanical properties of column foundations: Model tests.” Geotechnique, 50(6), 689–698.
reconstituted Bothkennar soil.” Geotechnique, 42(2), 289–301. Nash, D. F. T., Powell, J. J. M., and Lloyd, I. M. (1992a). “Initial investiga-
Balaam, N. P., and Booker, J. R. (1981). “Analysis of rigid rafts supported tions of the soft clay test site at Bothkennar.” Geotechnique, 42(2),
by granular piles.” Int. J. Numer. Anal. Methods Geomech., 5(4), 163–181.
379–403. Nash, D. F. T., Sills, G. C., and Davison, L. R. (1992b). “One-dimensional
Barksdale, R. D., and Bachus, R. C. (1983). “Design and construction of consolidation testing of the soft clay from Bothkennar.” Geotechnique,
stone columns.” Rep. No. FHWA/RD-83/026, National Information 42(2), 241–256.
Service, Springfield, VA. Paul, M. A., Peacock, J. D., and Wood, B. F. (1992). “The engineering geol-
Black, J. A., Sivakumar, V., and Bell, A. (2011). “The settlement perform- ogy of the Came clay of the national soft clay research site,
ance of stone column foundations.” Geotechnique, 61(11), 909–922. Bothkennar.” Geotechnique, 42(2), 183–198.
Brinkgreve, R. B. J., and Broere, W. (2006). PLAXIS 3D Foundation man- Priebe, H. J. (1976). “Evaluation of the settlement reduction of a founda-
ual, version 2, PLAXIS BV, Delft, Netherlands. tion improved by vibro-replacement.” Bautechnik, 5, 160–162 (in
Castro, J., and Sagaseta, C. (2009). “Consolidation around stone columns: German).
Influence of column deformation.” Int. J. Numer. Anal. Methods Priebe, H. J. (1995). “The design of vibro replacement.” Ground Eng.,
Geomech., 33(7), 851–877. 28(10), 31–37.
Castro, J., and Sagaseta, C. (2011). “Consolidation and deformation around Pulko, B., Majes, B., and Logar, J. (2011). “Geosynthetic-encased stone
stone columns: Numerical evaluation of analytical solutions.” Comput. columns: Analytical calculation model.” Geotext. Geomembr., 29(1),
Geotech., 38(3), 354–362. 29–39.

© ASCE 04016114-13 Int. J. Geomech.

Int. J. Geomech., 04016114


Rowe, P. W. (1962). “The stress-dilatancy relation for static equilibrium of soft clay deposit.” Proc., 2nd Int. Workshop on Geotechnics of Soft
an assembly of particles in contact.” Proc. R. Soc. London, Ser. A, Soils, CRC Press, Boca Raton, FL, 293–298.
269(1339), 500–527. Shahu, J. T., and Reddy, Y. R. (2011). “Clayey soil reinforced with stone
Schanz, T. (1998). “Zur modellierung des mechanischen verhaltens von rei- column group: Model tests and analyses.” J. Geotech. Eng., 10.1061
bungsmaterialien.” Thesis, Institut für Geotechnik, Univ. of Stuttgart, /(ASCE)GT.1943-5606.0000552, 1265–1274.
Stuttgart, Germany (in German). Wehr, W. (2004). “Stone columns: Single columns and group
Schanz, T., Vermeer, P. A., and Bonnier, P. G. (1999). “The hardening-soil behavior.” Proc., 5th Int. Conf. on Ground Improvement
model: Formulation and verification.” Beyond 2000 in Computational Techniques, CI-Premier Pte Limited, Singapore, 329–340.
Geotechnics: 10 Years of PLAXIS International, R. B. J. Brinkgreve, Wehr, W. (2006). “Stone columns—Group behaviour and influence of
ed., A. A. Balkema, Rotterdam, Netherlands, 281–290. footing flexibility.” Proc., 6th European Conf. on Numerical
Serridge, C. J., and Sarsby, R. W. (2008). “A review of field trials investi- Methods in Geotechnical Engineering, Taylor & Francis, London,
gating the performance of partial depth vibro stone columns in a deep 767–772.
Downloaded from ascelibrary.org by Ryerson University on 10/06/16. Copyright ASCE. For personal use only; all rights reserved.

© ASCE 04016114-14 Int. J. Geomech.

Int. J. Geomech., 04016114

You might also like