You are on page 1of 23

White, D. J. & Bolton, M. D. (2004). Géotechnique 54, No.

6, 375–397

Displacement and strain paths during plane-strain model pile installation


in sand
D. J. W H I T E * a n d M . D. B O LTO N *

The underlying mechanisms governing the behaviour of Les mécanismes sous-jacents gouvernant le comportement
displacement piles in sand are not well understood, lead- de piles de déplacement dans du sable ne sont pas bien
ing to unreliability in design methods. A series of plane- compris, ce qui nuit à la fiabilité des méthodes de
strain calibration chamber tests has been conducted in conception. Nous avons effectué une série d’essais de
order to quantify the penetration mechanism around the plan-déformation en chambres de calibrage afin de quan-
pile tip, and the response of the interface layer adjacent tifier le mécanisme de pénétration autour des extrémités
to the shaft during further penetration. A series of eight de piles ainsi que la réponse de la zone interface adja-
tests is reported, examining the influence of soil type, cente au tronc pendant la pénétration consécutive. Nous
initial state, pile breadth and the use of a driving shoe. A rapportons les résultats d’une série de huit tests, exami-
novel image-based deformation measurement technique nant l’influence du type de sol, l’état initial, la largeur de
has been used to observe the displacement and strain la pile et l’utilisation d’un sabot moteur. Nous utilisons
paths, which are found to be relatively independent of une nouvelle technique à imagerie pour mesurer la
soil type. The measured strain paths are similar to déformation et pour observer les chemins de déplacement
predictions made by the strain path method, and contrast et de déformation qui se sont révélés être relativement
sharply with assumptions implicit in cavity expansion indépendants du type de sol. Les chemins de déformation
solutions. An interface zone adjacent to the pile shaft mesurés sont similaires aux prévisions faites par la méth-
comprising fine broken soil particles was observed to ode de chemin de déformation et contrastent nettement
contract while shearing along the pile–soil interface. This avec les suppositions implicites dans les solutions d’ex-
mechanism offers an explanation for the degradation of pansion de cavité. Nous avons constaté qu’une zone inter-
shaft friction at a given soil horizon with increased pile face adjacente au tronc de la pile et constituée de fines
penetration (‘friction fatigue’), and a subsequent recovery particules de sol brisé se contractait tout en produisant
of capacity over time (‘set-up’). un cisaillement le long de l’interface pile-sol. Ce méca-
nisme offre une explication de la dégradation de la
friction du tronc à un horizon de sol donné avec pénétra-
KEYWORDS: bearing capacity; calcareous soils; friction; tion accrue de la pile (‘fatigue de pile’) et une récupéra-
model tests; piles; sands tion consécutive de la capacité dans le temps (‘reprise’).

INTRODUCTION tration test (CPT) data instead of soil properties for design.
The axial capacity of displacement piles in sand is arguably This approach bypasses the need to assume a penetration
the subject of greatest uncertainty in geotechnical engineer- mechanism. For example, by linking base resistance, qb , to
ing (Randolph et al., 1994). The most widely used technique CPT resistance, qc , the MTD Method (Jardine & Chow,
for the design of long offshore displacement piles is that 1996) can be used to predict base resistance in silica sand
proposed by the American Petroleum Institute (API, 1993). with a higher reliability than the API method (coefficient of
However, a large variation in Qpredicted /Qmeasured is exhibited variance (COV) in Qpredicted /Qmeasured ¼ 0.20 when measured
by this method when compared with a comprehensive data- against the database assembled by Chow, 1997). This im-
base of load tests (Chow, 1997). proved reliability is achieved by introducing a scale effect in
A number of approaches exist for predicting the base which qc is factored down by the ratio [1  0.5 log (D/
resistance of piles in sand. Analyses that use slip planes and Dcpt )]. The mechanistic origin of this scale effect is unclear,
bearing capacity theory to link friction angle with base resis- and indeed an opposite scale effect has been observed in
tance (e.g. Berezantzev et al., 1961) do not capture the trends small-scale centrifuge modelling (Klotz & Coop, 2001). A
observed in the field. Alternative methods range from wholly number of alternative reduction factors to convert qc to qb ,
empirical methods (e.g. API, 2000), based on soil type and based on different mechanistic hypotheses, have been sug-
relative density, to theoretical analyses based on cavity expan- gested (Winterkorn & Fang, 1975; Tejchman & Gwizdala,
sion (e.g. Randolph et al., 1994) or strain path methods 1979; Kraft, 1990; Lee & Salgado, 1999; Borghi et al.,
(Baligh, 1985). However, these methods also exhibit signifi- 2001; White & Bolton, 2004).
cant unreliability, and make contradictory assumptions regard- The observed unreliability in pile design methods suggests
ing the failure mechanism, indicating that the correct link that the governing behaviour is not well understood. There is
between soil properties and base resistance has not yet been no consensus for the mechanism by which a pile penetrates
identified. A recent prediction contest confirmed that pile sand or for the soil properties that govern this behaviour.
design methods remain highly unreliable (Jardine et al., 2001). This paper describes a series of plane-strain calibration
Improved reliability can be achieved by using cone pene- chamber tests in which the penetration mechanism of a
displacement pile is observed, allowing displacement and
Manuscript received 17 February 2003; revised manuscript accepted
strain paths to be measured. Previous attempts to quantify
14 April 2004. these soil movements (e.g. Robinsky & Morrison, 1964;
Discussion on this paper closes on 1 March 2005, for further Mikasa & Takada, 1973; Davidson et al., 1981; Allersma,
details see p. ii. 1988; Chong, 1988) have been hampered by poor measure-
* Department of Engineering, University of Cambridge, UK. ment resolution, the need to use artificial ‘soil’ material, and

375

Downloaded by [ University Of Western Australia] on [25/05/20]. Copyright © ICE Publishing, all rights reserved.
376 WHITE AND BOLTON
the inability of the apparatus to replicate field stress levels two. This has derived from the historic tendency for sands
around the pile tip. These disadvantages are overcome to be tested at stresses that are too low for the clearly
through the development of an image-based deformation defined common reference states of a linear critical state line
measurement system based on particle image velocimetry (CSL) and a normal compression line (NCL) to be reached
and close-range photogrammetry (White et al., 2003), and (e.g. Airey et al., 1988).
the construction of a reinforced plane-strain calibration Instead, in very general terms, crushable sands have been
chamber capable of revealing the soil surrounding the tip of characterised by high peak angles of friction, which decrease
a displacement pile while resisting the high penetration- with stress level and decay sharply with shear strain
induced stresses. It is noted, however, that plane-strain (Yasufuku & Hyde, 1995). Compression of crushable sands
geometry causes ground movements to extend further from leads to significant volume change due to particle breakage,
the pile than would occur in axisymmetry. For the case of which offsets the influence of dilation due to rearrangement.
an infinitely long pile installed in incompressible soil, lateral Uncrushable sands are characterised by a smaller variation
movements are inversely proportional to distance from the in angle of friction with stress level and very high volu-
pile, whereas in plane-strain they are constant. metric stiffness. Significant dilation occurs during shearing
of dense samples.
However, when taken to high stresses such as those
CALIBRATION CHAMBER induced by pile installation, sands that are considered crush-
Calibration chambers are widely used to study penetration able, as well as those not normally thought so, reveal similar
resistance (e.g. Houlsby & Hitchman, 1988; Yasufuku & patterns of behaviour. These patterns have much in common
Hyde, 1995; Salgado et al., 1997). The stresses and deforma- with critical state soil mechanics, and are often related to
tions around the tip of an advancing CPT or pile can be particle breakage (Coop, 1990). Two sands were selected for
correctly replicated by applying a surcharge pressure at the this research project, representing two extremes of mineral-
boundary of the chamber. In this research, a plane strain ogy and particle strength: Dog’s Bay carbonate sand (DBS)
chamber with observation windows has been constructed in and Leighton Buzzard Fraction B silica sand (LBS).
order to observe the soil deformation around an advancing Dog’s Bay sand is of biogenic origin with 88–94%
pile (Fig. 1). The front face of the box includes six viewing carbonate content (Houlsby et al., 1988), and consists of
windows through a 72.8 mm thick Perspex sheet. These view- broken shell fragments. The mechanical properties have been
ing windows allow observation of the advancing pile and the reported by Golightly & Hyde (1988), Coop (1990),
resulting soil movement. The Perspex sheet is supported by a Yasufuku & Hyde (1995) and Jovičič & Coop (1997).
grillage of deep I-section steel straps to minimise out-of- It should be noted that previous research has been con-
plane deflections. Sheets of 4 mm thick glass are placed on ducted using different gradings of Dog’s Bay sand. The
the inner faces of the box to reduce side friction. As the research described in this dissertation was carried out using
sand–glass interface friction angle is significantly lower than a natural grading of Dog’s Bay sand after passing through a
the soil friction angle, failure will occur at the interface 1.18 mm sieve to remove the gravel fraction. Index proper-
rather than within the soil body. This, coupled with the tight ties are shown in Table 1, and a photograph of typical grains
fit of the model pile between the front and rear faces of the is shown in Fig. 2(a).
chamber, ensures that the ground movement observed at the Leighton Buzzard sand (LBS) has been widely used in
window is equal to that present through the soil model. research for the past 50 years, and is a rounded silica sand
The internal width of the box is 1000 mm. Two breadths (Fig. 2(b)). Although no longer quarried from Leighton
of model pile were tested: 32.2 mm and 16.1 mm, corre- Buzzard, the material originates from the same geological
sponding to ratios of chamber width to pile breadth of 31 deposit, the Lower Greensand. Fraction B is the 0.6–
and 62 respectively. The model pile was jacked into the 1.18 mm sieve fraction, and is often referred to as 14/25.
chamber by a machine screw actuator driven by a stepper The mechanical behaviour of this fraction has been investi-
motor through a reduction gearbox. gated by Stroud (1971), Budhu (1979), Lee (1989) and
The model piles were machined from Duraluminium, and Schnaid (1990). Index properties of the sand used in this
fitted with 3 mm thick stainless steel base plates. A button research are shown in Table 1. For both sands, the maximum
load cell was mounted within the base plate with the active voids ratio was found using the procedure described in BS
face machined flush with the base. A narrow hole was 1377 (BSI, 1990). To avoid grain crushing, the minimum
drilled longitudinally through the pile to carry the load cell voids ratio was found by vibrating a 1 kg sample without the
wiring. After assembly of the chamber, the rear face of the use of a surcharge weight. The resulting values match
pile was machined to provide a tight fit across the width. closely with those previously published (e.g. Stroud, 1971;
The pile surface was machined in the vertical direction. The Golightly & Hyde, 1988).
resulting roughness was measured using a scanning white
light interferometer. The mean roughness, RA , measured at
10 locations on the pile surface was found to be 0.326 m TEST PROCEDURE
parallel to machining, that is, in the vertical direction during A series of eight tests is reported, each using dry sand
installation, and 1.424 m perpendicular to machining. These (Table 2). The tests are identified by a code comprising the
values compare closely with other reported model pile tests, test number, sand type (DBS: Dog’s Bay sand, LBS: Leight-
which typically use piles of roughness 1–2 m (Bruno, on Buzzard sand) and relative density. The suffixes ‘shoe’
1999; Klotz, 2000). The tip of the pile was fitted with a and ‘narrow’ denote the tests with a pile shoe and half-width
driving shoe for test T7-DBS-45-shoe. The driving shoe pile respectively. In each case, the model pile was jacked
consisted of the standard base plate, extended to overhang monotonically from the surface of the sample. During the
either side of the pile by 4 mm. first six tests, the initial state of the sand was varied and
both silica and carbonate sands were used. Two additional
tests were conducted: one test using the narrower pile
TEST SANDS (16.1 mm breadth), and one test in which the broad pile was
The mechanical behaviour of sands is usually considered fitted with a driving shoe.
to depend on whether or not the particles are crushable, with The sand sample was prepared by pluviation in air, with
existing literature making a sharp distinction between the drop height and flow rate varied to achieve the desired initial

Downloaded by [ University Of Western Australia] on [25/05/20]. Copyright © ICE Publishing, all rights reserved.
PLANE-STRAIN MODEL PILE INSTALLATION IN SAND 377

Actuator not shown

Internal width
1000 mm

Sand depth
745 mm

(a)

Actuator not shown

Reaction frame

Rear panel assembly

Surcharge bags

Front panel assembly

Window assembly

Window 3
Window 4

(b)

Fig. 1. Schematic arrangement of calibration chamber: (a) front and side projections of assembled chamber;
(b) exploded view of chamber components

Table 1. Index properties of test sands


Sand Mineralogy Gs D50 : mm emax emin
Dog’s Bay sand Calcium carbonate 2.75* 0.44 1.87 0.98
Leighton Buzzard Fraction B Silica 2.65† 0.84 0.80 0.51
*From Golightly (1989).
†From Tan (1990).

Downloaded by [ University Of Western Australia] on [25/05/20]. Copyright © ICE Publishing, all rights reserved.
378 WHITE AND BOLTON

Fig. 2. SEM image of each test sand: (a) Dog’s Bay carbonate sand (SEM image from Bowman et
al., 2001); (b) Leighton Buzzard Fraction B silica sand (SEM image from Sentenac et al., 2001)

state. Each model was poured in lifts of approximately The model pile was jacked into the chamber at a rate of
50 mm with density checks carried out after each lift. The 1 mm/min. Digital cameras were used to record images of
relative density of lifts within a single model fell within a the soil and pile at regular intervals. A button-type load cell
range of 2%. A surcharge of 100 kPa was applied to the installed in the tip of the model pile recorded base resis-
top surface of the sand, and reduced to 50 kPa. tance, qb , throughout installation. Feedback from the stepper

Downloaded by [ University Of Western Australia] on [25/05/20]. Copyright © ICE Publishing, all rights reserved.
PLANE-STRAIN MODEL PILE INSTALLATION IN SAND 379
Table 2. Calibration chamber test series
Test code Sand type Initial voids ratio, e0 Relative density: % Pile breadth: mm Pile tip configuration
T1-DBS-64 DBS 1.30 64 32.2 Flat ended
T2-DBS-44 DBS 1.48 44 32.2 Flat ended
T3-DBS-71 DBS 1.24 71 32.2 Flat ended
T4-LBS-34 LBS 0.70 34 32.2 Flat ended
T5-LBS-55 LBS 0.64 55 32.2 Flat ended
T6-DBS-84 DBS 1.12 84 32.2 Flat ended
T7-DBS-45-shoe DBS 1.47 45 32.2 Driving shoe
T8-DBS-46-narrow DBS 1.46 46 16.1 Flat ended

motor was used to monitor the pile tip depth, with cross- 0
checks made from the camera images.
A set of image control points was installed within each 50
T4-LBS-34 T5-LBS-55
porthole in the viewing window. These control points were
laser-printed onto overhead transparency sheets and meas- 100

Pile tip depth below sand surface, z: mm


ured using a calibrated photogrammetric target to establish
their exact locations, eliminating errors due to inaccurate 150
laser printing. These control points were used to establish
the photogrammetric transformation parameters required to 200 T6-DBS-84
convert image-space PIV measurements of soil movement T3-DBS-71
250 T7-DBS-45-shoe
into object-space coordinates (White et al., 2003). The over- T1-DBS-64
head transparency sheets were installed between the Perspex
300
window and the glass sheets. An appropriate refractive
correction was included in the transformation to account for
350
the offset distance between the observed plane of soil and
the plane of the control points. 400

450 T2-DBS-44
T8-DBS-46-narrow
TEST RESULTS 500
The primary source of data in each test was the images 0 2 4 6 8 10
captured through the viewing portholes. Soil displacement Base resistance, qb: MPa
and strain paths around the advancing pile were measured
from these images. In addition, the load cell in the pile base 0
indicated the qb –depth profile. Fig. 3 shows the variation of
50
qb with pile tip depth for each test. Within each test there is
a general trend of sharply increasing base resistance during
100
Pile tip depth below sand surface, z: mm

the first 30–50 mm of penetration, during which the pile can


be considered as a shallow foundation, followed by a more T5-LBS-55
150
steady response. An approximately constant value of qb and
therefore a steady-state penetration mechanism was reached 200
during the tests on DBS.
During the LBS tests, a constant value of qb was not 250
reached. During the loose test (T4-LBS-34) an increase in
qb of 32% was recorded between depths of 100 mm and 300 T4-LBS-34
440 mm. The rising gradient of the qb –depth curves for the
LBS tests indicates that a steady state is not being ap- 350
proached. This trend could derive from an increasing influ-
ence of the chamber base with pile tip depth, as the 400
influence of a remote rigid boundary would be larger in the
stiff silica sand than in the compressible carbonate sand. 450
Klotz & Coop (2001) also observed boundary effects when
installing model piles in LBS, whereas tests in DBS using 500
0 5 10 15 20 25 30 35
the same chamber appeared unaffected. Test T5-LBS-55 was Base resistance, qb: MPa
halted at a depth of 301 mm, as the base resistance was
approaching the maximum design load of the chamber. Fig. 3. Base resistance, qb , plotted against pile tip depth during
The steady-state values of base resistance reached during installation
the DBS tests allow the influence of initial soil state to be
examined. A trend of increasing qb with initial relative
density is evident from Fig. 4. If it is assumed that the depth of 100 mm the base resistance in tests of comparable
ambient vertical stress in the region of steady-state penetra- relative density differs by a factor of 6–8.
tion is 50 kPa, the bearing capacity factor, Nq , varies in the
range 36–116 between relative densities of 44% (T8-
DBS-46-narrow, qb ¼ 1.82 MPa) and 84% (T6-DBS-84, qb PHOTOGRAPHIC DATA
¼ 5.8 MPa). As would be expected, significantly higher base The technique used to measure soil displacement from
resistance is measured in LBS than in DBS. At a pile tip digital images is described by White et al. (2003). This

Downloaded by [ University Of Western Australia] on [25/05/20]. Copyright © ICE Publishing, all rights reserved.
380 WHITE AND BOLTON
6 the above measurements from each test. Selected measure-
Steady-state base resistance, qb: MPa
T6-DBS-84 ments are presented in order to establish:
5 (a) the influence of the chamber boundaries
T3-DBS-71 (b) the mechanism of penetration, by which base resistance
4 is created
(c) the soil movement adjacent to the pile shaft, which
T1-DBS-64
strongly influences shaft friction.
3
Images were captured at pile tip displacement increments
T2-DBS-44 of 1–2 mm. A typical image, from test T2-DBS-44 captured
2 at a pile embedment of 300 mm, is shown in Fig. 5, overlain
T8-DBS-46-narrow by a mesh of 20 3 20 pixel PIV test patches. Each PIV
1 patch is nominally 2.5 mm 3 2.5 mm in size, covering only
a few individual grains. As a result, the ‘gauge length’
between adjacent displacement measurements, from which
0 strains are calculated, is too small for a smooth strain path
40 45 50 55 60 65 70 75 80 85
Initial relative density, ID to be expected. In the same manner, a smooth stress–strain
curve would not be expected from an element test of
Fig. 4. Variation of steady-state base resistance with relative dimensions 2.5 3 2.5 mm.
density

DISPLACEMENT PATHS DURING PILE INSTALLATION


technique combines particle image velocimetry (PIV) and The most straightforward illustration of the penetration
close range photogrammetry to achieve significantly higher mechanism is the displacement field around the pile tip. Fig.
accuracy and precision than previous methods used in 6 shows the displacement field found from comparison of
geotechnical modelling, without requiring target markers to the PIV mesh shown in Fig. 5 with a subsequent image,
be installed in the observed soil. captured after 1.5 mm of pile tip movement. This displace-
The following measurements were deduced from the ment field consists of 3901 vectors, which have been en-
captured images (White, 2002): larged to reveal the deformation pattern. Fig. 7 shows the
(a) soil displacement during pile installation magnitude of the displacement vectors in Fig. 6, indicating
(b) soil strain paths during pile installation the extent of the displacement field. Although the deforma-
(c) streamlines of soil flow tion consists primarily of downward movement below the
(d) soil strain post-installation (i.e. as would be the case pile, horizontal displacements of 25 m are detected in the
prior to loading of pile) far field, at an offset of 160 mm ( 5B) from the pile
(e) soil movement adjacent to pile shaft. centreline.
There is no evidence of a bearing capacity-type mechan-
Limitations of space prevent this paper from containing all ism in which the soil flows along streamlines curving from

200

400
Vertical pixel coordinate, u

600

800

1000

0 400 800 1200 1600


Horizontal pixel coordinate, v

Fig. 5. Typical PIV mesh

Downloaded by [ University Of Western Australia] on [25/05/20]. Copyright © ICE Publishing, all rights reserved.
PLANE-STRAIN MODEL PILE INSTALLATION IN SAND 381
1 mm displacement 5 mm displacement

120

100
Vertical coordinate: mm

80

60

40

20

0
0 50 100 150 200

Horizontal coordinate: mm

Fig. 6. Displacement field around pile tip

0·025
120 1·5

0·025 1·4
1·3
100 1·1
1·2

0·025
Vertical coordinate: mm

0·1
80 0·8 0·9
0·075 0·4
0·7 0·6
0·5
0·025 0·4
60
0·025
0·05
0·4
0·1
40
0·1
0·05 0·3
0·075
0·1 0·25
20 0·2
0·05
0·15
0·1
0·05
0
0 50 100 150 200
Horizontal coordinate: mm

Fig. 7. Contours of displacement field magnitude

below the pile tip around to the upward direction on either which all components are assumed to vary only with the
side of the shaft. The penetration mechanism is more radial coordinate.
comparable to cavity expansion, with the displacement vec- Figure 8 shows the full displacement trajectories of two
tors radiating from the pile tip downwards and to the side. soil elements tracked through tests T2-DBS-44 and T4-LBS-
However, although the contours of displacement follow a 34. The coordinate origin is located on the centreline of the
circumferential path immediately below the pile, these con- pile and level with the pile tip at the end of installation. As
tours return to the pile shoulder. This sharp variation in the pile approaches, the movement is generally downwards,
instantaneous velocity with inclination from the vertical is with the soil element trajectory curving towards the horizon-
in contrast to a cavity expansion model for penetration, in tal as the pile passes. In the case of LBS, the final part of

Downloaded by [ University Of Western Australia] on [25/05/20]. Copyright © ICE Publishing, all rights reserved.
382 WHITE AND BOLTON
Pile centreline
100·7
150
100·6

140
100·5

h⫽0
130 100·4

120 100·3

100·2
110

100·1
100
100
20 10 0 21·8 21·7 21·6 21·5 21·4 21·3 21·2 21·1 21 20·9 20·8

(a)
Pile centreline
145 104·9

140
104·8

135
104·7
130
104·6 h⫽0
125
104·5
120

104·4
115

110 104·3

105 104·2

100
25 20 15 10 5 0 ⫺5 20·6 20·5 20·4 20·3 20·2 20·1 20 19·9 19·8 19·7

(b)

Fig. 8. Soil element trajectories during pile installation (coordinates in mm): (a) T2-DBS-44; (b) T4-LBS-34

the element trajectory is upwards, although the net move-


ment is downwards, in contrast to bearing capacity mechan-
1·2
Net displacement ratio, 2∆z/B, 2∆x/B

isms, which rely on soil heave to accommodate the pile T2-DBS-44, 2∆z/B
volume. Greater net downward movement and less net lateral
movement is observed in DBS than in LBS. 1·0
A notable feature of Fig. 8 is the ‘tail’ at the end of each T2-DBS-44, 2∆x/B T4-LBS-34, 2∆x/B
trajectory. After the pile tip has passed the soil element (i.e. 0·8
h . 0, where h is the vertical distance of a soil element
above the pile tip), the soil relaxes back towards the pile 0·6
shaft. This ‘tail’ in the displacement trajectory is examined
more closely in connection with shaft friction later in this 0·4 T4-LBS-34, 2∆z/B ∆x
paper.
∆z
0·2
x0 ⫽ 1·15B B
STEADY-STATE DEFORMATION
0
In Fig. 9 the displacement trajectories for the column of 280 300 320 340 360 380 400 420
soil elements visible in window 4 and initially located at an Initial soil element depth, z: mm
offset of 1.15B from the pile centreline in tests T2-DBS-44
and T4-LBS-34 have been normalised by the half-breadth of Fig. 9. Net displacement of a single column of soil elements

Downloaded by [ University Of Western Australia] on [25/05/20]. Copyright © ICE Publishing, all rights reserved.
PLANE-STRAIN MODEL PILE INSTALLATION IN SAND 383
1·0
the model pile. The horizontal and vertical displacement Symmetry boundary condition
ratios are defined as the net horizontal movement, ˜x, and 0·9
net downward movement, ˜z, during installation of the pile

Horizontal displacement ratio, 2∆x/B


0·8 T4-LBS-34
(until h ¼ 0), normalised by the pile half-breadth, B/2. Fig. T5-LBS-55
0·7
9 demonstrates the steadiness of the deformation pattern;
there is no systematic trend for 2˜x/B or 2˜z/B to increase 0·6

or decrease with depth as the pile passes the viewing 0·5


window. 0·4
Figure 9 does reveal some random variation in displace- T3-DBS-71
T8-DBS-46-narrow T2-DBS-44
ment ratio of approximately 0.03, which is equal to a 0·3
T7-DBS-45-shoe
displacement of 400 m—approximately one particle dia- 0·2 Chamber boundary for all tests
except T8-DBS-46-narrow T6-DBS-84
meter. Random noise of this magnitude is to be expected, 0·1
considering that the PIV patch size is equivalent to a few 0
particle diameters. It is highly conceivable that while being 35 30 25 20 15 10 5 0
displaced by 55 mm (.125 particle diameters) during instal- Initial offset from pile centreline, 2x0/B
(a)
lation of the pile, as shown in Fig. 8, adjacent groups of four
0
to seven particles may become separated from their neigh- T5-LBS-55
T4-LBS-34
bours by a particle diameter. The assumption of continuum T8-DBS-46-narrow
⫺1
behaviour therefore breaks down, and adjacent pairs of soil

Vertical displacement ratio, 2∆z/B


elements cannot be used to generate continuum strain paths T2-DBS-44
T3-DBS-71
for the bulk material. ⫺2 T6-DBS-84
T7-DBS-45-shoe
The steady-state deformation pattern shown in Fig. 9
shows that displacement trajectories for a given value of h ⫺3
will be the same, independent of the original depth of the
soil element. Therefore all trajectories from a single column ⫺4
of soil elements can be combined to provide the deformation
pattern over a greater vertical extent below the pile tip than ⫺5
is visible at any given instant.
The viewing window only allows a vertical extent of ⫺6
160 mm (5B) to be simultaneously observed. However, by 15 10 5 0
combining the displacement data during installation found Initial offset from pile centreline, 2x0/B
(b)
from all soil elements originally located within a single
vertical column, the penetration mechanism can be inferred Fig. 10. (a) Horizontal and (b) vertical displacement during pile
over a greater vertical extent. Furthermore, this method of installation
combining the data from multiple patches allows strain
measurements representative of the continuum behaviour to
be obtained. more conservative ratio of chamber to pile size than axisym-
This transformation is equivalent to describing every soil metric geometry (e.g. Bolton et al., 1999) if the displace-
element trajectory using a reference frame with the origin ment field around the pile is to be entirely uninfluenced by
located at the tip of the pile (on the centreline). The the chamber boundaries. The horizontal displacement ratio
penetration mechanism is then viewed as the steady flow of close to the pile shaft during test T8-DBS-46-narrow con-
soil around a stationary pile. The coordinate axes consist of verges towards the data from tests conducted using the larger
the horizontal offset from the centreline, x, and the vertical pile, suggesting that although the far-field behaviour is
distance above the tip of the pile, h (h is positive upwards). affected by the end wall, the influence in the near field is
To make clear the distinction between x–y plots of minimal.
displacement data from individual patches (as shown in Figs The curves of displacement ratio shown in Fig. 10 divide
6–9), and x–h plots that have been assembled by combining by soil type. The zone of horizontal deformation is concen-
the data from multiple patches in the column-wise manner trated closer to the pile shaft in compressible DBS than in
described above, all x–h plots (Figs 16–19) are presented as LBS, and DBS is pushed downwards almost twice as far as
mirror images with the pile on the left-hand side. In this the LBS. For both soils the zone of downward displacement
mode of presentation the x-axis is positive in the left-to-right is concentrated closely around the pile shaft, whereas the
direction. horizontal displacement decays slowly with offset distance.
For each test a network of triangular strain elements was
constructed from the mesh of soil elements tracked by PIV.
NET SOIL MOVEMENT DURING INSTALLATION The procedure described in Appendix 1 was used to decom-
Figure 10 shows the variation in horizontal and vertical pose the deformation into strain (or stretch) and rotation.
displacement ratio with initial offset from the pile centreline, Fig. 11 shows a zone of very high rotation (.208) close to
2x0 /B, for tests T2-DBS-44 to T8-DBS-46-narrow, using data the pile (2x0 /B , 2), justifying the use of a finite rotation
both from window 4, through which the pile can be ob- strain calculation. The rigid body rotation measured in DBS
served, and from window 3, through which the far-field is approximately twice that measured in LBS at a compar-
behaviour (x0 ¼ 6.5B–12B) can be seen. able initial offset.
It is evident from Fig. 10 that the end wall of the chamber
has an influence on the displacement field around the pile.
All curves for horizontal displacement ratio tend towards SOIL STRAIN PATHS DURING PILE INSTALLATION
zero at 2x0 /B  31, which corresponds to the end wall The calculation procedure described in Appendix 1 con-
position. The exception is test T8-DBS-46-narrow. In this verts the measured displacement trajectories into the strain
case, the end wall is located at 2x0 /B ¼ 62, towards which paths induced by pile installation. In order to present the
the horizontal displacement tends to zero. This observation evolution of strain within a soil element as it flows towards
demonstrates that plane-strain geometry requires a much and beyond the pile tip, strain paths are presented as shown

Downloaded by [ University Of Western Australia] on [25/05/20]. Copyright © ICE Publishing, all rights reserved.
384 WHITE AND BOLTON
60 shaft. This zone extends horizontally by only 2–3 mm and
therefore cannot contain sufficient PIV patches for strain
50 paths to be calculated. Other evidence of the volume change
Soil element rotation (clockwise

behaviour within the interface zone based on post-test


sampling is discussed later.
positive): degrees

40 T2-DBS-44
T3-DBS-71
T6-DBS-84 The very near field behaviour is indicated by the strain
T7-DBS-45-shoe
T8-DBS-46-narrow
paths of soil elements originally located at 2x0 /B ¼ 1 and
30
2x0 /B ¼ 2. These strain paths reach values of natural direct
strain greater than 50% in compression and 200% in exten-
20 sion, which far exceed the range of conventional laboratory
T4-LBS-34
T5-LBS-55 element testing. The shape of the strain path in the near
10 field is similar for both soil types. Below the pile tip the
principal strain directions are approximately vertical and
0
horizontal, as required for symmetry.
15 10 5 0 A reversal of the vertical strain rate from compression to
Initial offset from pile centreline, 2x0/B
extension occurs at 2h/B  3, followed by a reversal in
the horizontal strain rate from extension to compression at
Fig. 11. Soil element rotation during pile installation
2h/B  1, accompanied by significant rotation. On passing
the pile tip, the soil elements lie in horizontal extension
in Fig. 12, having been decomposed from rigid body rota- relative to their initial state, albeit having rotated by as much
tion. Natural (or logarithmic) strains are used, as is conven- as 458. Although the general shapes of the strain paths in
tional for large-strain behaviour, with compression positive. DBS and LBS are alike, the strain levels in the very near
The position of the soil element is described by the h- field are 30–50% higher in DBS for a given value of 2h/B.
coordinate, normalised by B/2. The strain quantities included The near-field behaviour is indicated by the strain paths of
on the plot are the cumulative horizontal and vertical strain, soil elements originally located at 2x0 /B ¼ 3, 4 and 6. Once
defined within the reference frame of the soil element. Also again, similar trends are observed in DBS and LBS. Within
shown are the maximum and minimum principal strains. The this range of offset the absolute values of strain are of
evolution of these strain quantities can be regarded as the comparable magnitude in both sands. Compared with the
projection of a Mohr’s circle of strain along the shear strain very near field, the change in direction of strain rate occurs
axis. This construction indicates the maximum shear strain earlier, and at the same point in both the horizontal and
by the width of the envelope created by I and II . It should vertical directions. This is followed by strain sufficient to
be noted that, for very high values of natural strain, the exceed the initial cycle. Both the horizontal and vertical
usual construction of a Mohr’s ‘circle’ does not apply. For strain paths return to and cross the strain axis. The final
example, the mean value of the principal strains is not equal state is one of horizontal compression and vertical exten-
to the mean value of the vertical and horizontal strains, as is sion.
the case for a small-strain formulation. Shear strain continues to accumulate after reversal of the
Figures 13 and 14 show strain paths for tests T2-DBS-44 horizontal and vertical strain direction. The point at which
and T4-LBS-34 from six different initial horizontal offsets horizontal and vertical strain are equal (and close to zero)
from the pile centreline, corresponding to 2x0 /B  1, 2, 3, 4, corresponds closely with the point of maximum shear strain
6 and 10. The variation of rigid body rotation with 2h/B for (i.e. maximum principal strain difference), indicating a state
each initial offset is also shown. The volumetric strain paths of almost pure shear.
are shown separately in Fig. 15. Volumetric strain is defined The far-field behaviour is indicated by the strain paths of
in the conventional manner as the change in volume ex- soil elements originally located at 2x0 /B  10. At this large
pressed as a percentage of the original volume. offset from the pile shaft, the shape of the strain path differs
These six strain paths are divided into three zones of slightly between DBS and LBS. In DBS the trend of initial
behaviour—‘very near’, ‘near’ and ‘far’ field—and discussed horizontal extension followed by compression seen in the
separately below. An additional zone of behaviour, referred near field is again observed. In LBS the horizontal strain
to as the interface zone, lies immediately adjacent to the pile increases monotonically in compression. Also, the absolute

Plane of Upward movement of


projection soil relative to the pile tip
γ/2 ε

εI
εv

εv

εII εI ε

2h/B
εh

εII
εh

Fig. 12. Illustration of strain path by evolving Mohr’s circle of strain

Downloaded by [ University Of Western Australia] on [25/05/20]. Copyright © ICE Publishing, all rights reserved.
PLANE-STRAIN MODEL PILE INSTALLATION IN SAND 385
2x0/B ⫽ 1·06, 2xf /B ⫽ 1·84 2x0/B ⫽ 4·15, 2xf /B ⫽ 4·69
50 εI 15
εv v εI
0 10

⫺50 5
Strain, ε: %

Strain, ε: %
εhh
εhh
⫺100 0 εv v

⫺150 ⫺5

εII
⫺200 ⫺10 εII

⫺250 ⫺15
⫺25 ⫺20 ⫺15 ⫺10 ⫺5 0 5 ⫺25 ⫺20 ⫺15 ⫺10 ⫺5 0 5
Position relative to pile tip, 2h/B Position relative to pile tip, 2h/B

2x0/B ⫽ 2·08, 2xf /B ⫽ 2·73 2x0/B ⫽ 6·08, 2xf /B ⫽ 6·55


30 8
εI
20 6 εI
10 4
εv v

Strain, ε: %
0 εhh
Strain, ε: %

2
⫺10 εhh
0
⫺20 εv v
⫺2
⫺30

⫺4 εII
⫺40 εII
⫺50 ⫺6
⫺25 ⫺20 ⫺15 ⫺10 ⫺5 0 5 ⫺25 ⫺20 ⫺15 ⫺10 ⫺5 0 5
Position relative to pile tip, 2h/B Position relative to pile tip, 2h/B
2x0/B ⫽ 3·1, 2xf /B ⫽ 3·69 2x0/B ⫽ 9·92, 2xf /B ⫽ 10·3
20 4

15 εI
3 εI
10
2
εhh
5
Strain, ε: %
Strain, ε: %

εhh 1
0 εv v
0
⫺5
εv v
⫺10 ⫺1

⫺15 ⫺2 εII
εII
⫺20 ⫺3
⫺25 ⫺20 ⫺15 ⫺10 ⫺5 0 5
⫺25 ⫺20 ⫺15 ⫺10 ⫺5 0 5
Position relative to pile tip, 2h/B
Position relative to pile tip, 2h/B
2x0 /B
40
Rigid body rotation (clockwise positive):

1·06
35
εI, εII: maximum
B 30
and minimum
principal strains
25
degrees

εv v, εhh: vertical 20
and horizontal xf h 2·08
strains (within 15
reference frame 10 3·10
of element)
5 4·15
6·08
All strains are x0 0 9·92
natural (logarithmic)
Compression is positive ⫺5
⫺25 ⫺20 ⫺15 ⫺10 ⫺5 0 5
Position relative to pile tip, 2h/B

Fig. 13. Strain and rotation paths during pile installation: T2-DBS-44

strain levels in DBS are typically 50% lower than in LBS. tonically increasing compression with increasing 2h/B in
This is in agreement with Fig. 10, in which the deformation both sands, with greater compression observed for DBS
is localised closer to the pile tip in DBS than LBS. (Fig. 15). Immediately below the pile tip (4 , 2h/B ,
The volumetric strain paths in DBS and LBS show mono- 1), dilation is evident in the very near field (2x0 /B , 2),

Downloaded by [ University Of Western Australia] on [25/05/20]. Copyright © ICE Publishing, all rights reserved.
386 WHITE AND BOLTON
2x0/B ⫽ 0·97, 2xf /B ⫽ 1·76 2x0/B ⫽ 3·92, 2xf /B ⫽ 4·67
50 10
εI
εI
εv v
0 5
Strain, ε: % εhh

⫺50

Strain, ε: %
εhh 0
εv v

⫺100 ⫺5
εII
εII
⫺150 ⫺10

⫺200 ⫺15
⫺25 ⫺20 ⫺15 ⫺10 ⫺5 0 5
⫺25 ⫺20 ⫺15 ⫺10 ⫺5 0 5
Position relative to pile tip, 2h/B Position relative to pile tip, 2h/B
2x0/B ⫽ 1·99, 2xf /B ⫽ 2·8 2x0/B ⫽ 5·92, 2xf /B ⫽ 6·6
20 8
εI
10 6
εI

εv v 4
0 εhh

Strain, ε: %
Strain, ε: %

εhh
2
⫺10
0
⫺20 εv v
εII ⫺2
⫺30 εII
⫺4

⫺40 ⫺6
⫺25 ⫺20 ⫺15 ⫺10 ⫺5 0 5 ⫺25 ⫺20 ⫺15 ⫺10 ⫺5 0 5
Position relative to pile tip, 2h/B Position relative to pile tip, 2h/B

2x0/B ⫽ 3·01, 2xf /B ⫽ 3·78 2x0/B ⫽ 10·03, 2xf /B ⫽ 10·55


15 5
εI
4
10 εhh
εI
3
5
2
Strain, ε: %
Strain, ε: %

εhh
0 εv v 1

0
⫺5

εII ⫺1
⫺10 εv v
⫺2 εII
⫺15 ⫺3
⫺25 ⫺20 ⫺15 ⫺10 ⫺5 0 5 ⫺25 ⫺20 ⫺15 ⫺10 ⫺5 0 5
Position relative to pile tip, 2h/B Position relative to pile tip, 2h/B
2x0/B
30
Rigid body rotation (clockwise positive):

0·97
εI, εII: maximum 25
B
and minimum
principal strains 20

εv v, εhh: vertical
degrees

15
and horizontal xf h
1·99
strains (within 10
reference frame
of element) 5 3·01
3·92
All strains are 5·92
x0 0 10·03
natural (logarithmic)
Compression is positive ⫺5
⫺25 ⫺20 ⫺15 ⫺10 ⫺5 0 5
Position relative to pile tip, 2h/B

Fig. 14. Strain and rotation paths during pile installation: T4-LBS-34

followed by recompression as the soil element moves around ¼ 0.97. The end points of the volumetric strain paths
the pile shoulder (1 , 2h/B , 1). In DBS the soil remains indicate the horizontal variation in density adjacent to the
in net compression for h . 0 despite this dilation cycle, pile shaft (excluding the interface layer, which is discussed
whereas in LBS a net dilation of 1% is recorded for 2x0 /B later). In DBS, despite the dilation around the pile shoulder,

Downloaded by [ University Of Western Australia] on [25/05/20]. Copyright © ICE Publishing, all rights reserved.
PLANE-STRAIN MODEL PILE INSTALLATION IN SAND 387
12
2x0 /B 0
10
1·06
Volumetric strain: %

Vertical position relative to pile tip, 2h/B


8
⫺2
Horizontal compression,
6 vertical extension
2·08

4
3·10 ⫺4
4·15
6·08
2
9·92 ⫺6
0
⫺25 ⫺20 ⫺15 ⫺10 ⫺5 0 5
Position relative to pile tip, 2h/B
⫺8
(a)

4
10·03 3·01 5·92 3·92 ⫺10 Horizontal extension,
2 vertical compression
1·99
0 0 2 4 6 8 10
Volumetric strain: %

0·97 Horizontal position relative to pile tip, 2x/B


⫺2
(a)
⫺4

⫺6
0
⫺8

⫺10
Vertical position relative to pile tip, 2h/B

Label indicates 2x0 /B


⫺2
⫺12 Horizontal compression,
⫺25 ⫺20 ⫺15 ⫺10 ⫺5 0 5 vertical extension
Position relative to pile tip, 2h/B
(b) ⫺4

Fig. 15. Volumetric strain paths: (a) T2-DBS-44; (b) T4-LBS-34


⫺6

there is a continual decrease in density with increasing


offset. In LBS, the strong dilation in the near field creates a ⫺8
zone of soil close to the shaft that is less dense than the
more distant soil.
Horizontal extension,
⫺10
vertical compression

STREAMLINES OF SOIL FLOW 0 2 4 6 8 10


The two strain reversal points corresponding to the maxi- Horizontal position relative to pile tip, 2x/B
mum of v and minimum of h shown in Figs 13 and 14 are (b)
superimposed on the streamlines of soil flow in Fig. 16 for Strain direction transitions
the full range of initial horizontal offset. These points are Transition from vertical compression to extension
approximately coincident on each streamline, revealing two Transition from horizontal extension to compression
distinct zones of deformation. Below the pile and within a
region bounded by a line inclined downwards at approxi- Fig. 16. Streamlines of soil flow and strain rate reversal points:
mately 558 from the horizontal is a zone of soil that is (a) T2-DBS-44; (b) T4-LBS-34
undergoing vertical compression and horizontal extension.
Above this region the strain rates are reversed. Close to the
mately equal to a short reverse excursion along the strain
pile the strain reversal points do not coincide, and a short
paths in Figs 13 and 14 by a distance equal to 2˜h/B ¼ 0.1.
phase of both vertical and horizontal extension is observed.
The spatial distribution of installation-induced strains has
This corresponds to the zone of dilation evident in the
a strong influence on the settlement response of a displace-
volumetric strain paths.
ment pile under working load. The base stiffness of a pile
arises from integration of the stiffness contributions provided
by the deforming soil around the pile: if this soil is to be
POST-INSTALLATION STRAIN DISTRIBUTION modelled, appropriate stiffness values must be selected. A
The strain paths presented in Figs 13 and 14 can be bored pile can be ‘wished in place’ and analysed, based on
assembled as the strain distribution within the soil surround- the assumption that installation-induced strains are small and
ing the pile after installation. This spatial distribution of negligible. In contrast, the strain paths shown previously
strain also represents the initial conditions prior to applica- indicate that the condition of the soil surrounding a displace-
tion of a working load to the pile if the strain associated ment pile is significantly changed by the installation proce-
with removal of the installation force and during any interim dure. Soil stiffness is influenced by strain level (e.g. Jardine
creep period is ignored. The pile tip rebound during unload- et al., 1984) and recent loading direction (Atkinson et al.,
ing of a field pile might typically be equal to 1 or 2 mm. 1990). These influences can be quantified from the strain
This would induce strains in the surrounding soil approxi- distribution and the streamlines of flow.

Downloaded by [ University Of Western Australia] on [25/05/20]. Copyright © ICE Publishing, all rights reserved.
388 WHITE AND BOLTON
Figures 17 and 18 show the spatial distribution of strain test T8-DBS-46-narrow is shown, to allow the influence of
surrounding the base of the pile in tests T2-DBS-44 and T4- the ratio of chamber to pile size to be examined (Fig. 19).
LBS-34. These are the tests for which strain paths were To provide clarity in areas of high strain gradient, the
previously presented. In addition, the strain distribution from contour interval is varied from 0.5% in the far field to 25%

⫺100

⫺1·5
⫺50
⫺20
⫺10
⫺5
1·5

0·5
30
15
3

6
3

0
2 2

⫺0
·5

2·5
0 0

3
⫺10
0
⫺2 ⫺2
⫺5
5

1
1015
00
0

2
⫺1

⫺8

20


5 ⫺0·5

5
5
2

25 0
3
⫺4 ⫺4
⫺25 20

0
⫺305

⫺2 1 1·5 ⫺0·5

4
Vertical position relative to pile tip, 2h/B

Vertical position relative to pile tip, 2h/B


⫺6
0·5 0

⫺6 ⫺6

1

⫺0


5

5
·5
⫺10

2·5
0
⫺3


⫺8 ⫺8 ⫺8 10

3
1

2
⫺ 8

0·5
6 0
⫺2

6
5
⫺ 5
⫺10 ⫺10
⫺1·5

4

4

⫺12 ⫺12

0·5
⫺3
⫺0·5

1
3

1·5
0

⫺14 ⫺2·5 ⫺14

5

2
⫺2

⫺1

1·5
⫺1·5
⫺16 ⫺16
0

2
1·5 0·5
⫺0·5

⫺18 1
1
⫺18 1

⫺0·5 1 0·5
0

0
0·5
⫺20 ⫺20 0·5
0 ⫺0·5 2 4 6 8 10 0 2 4 6 8 10

Horizontal position relative to pile tip, 2x/B Horizontal position relative to pile tip, 2x/B
(a) (b)
12
10
6
5

4
4

2 2
2·5
5

12
1·5

0
14

0
1

⫽ 32%
3
16

1·5
2·5

5
3·5
75

40
50

0·5
2

⫺2 ⫺2
12

4
3
30

150
2
10

⫺4
0

1·5

⫺4
10
8

1·5
25

3
Vertical position relative to pile tip, 2h/B

Vertical position relative to pile tip, 2h/B

40
20
⫺6 30
9
⫺6 1
5

1
25
5
7

16

14
10

20
⫺8 16 ⫺8 2
6

12
14 1·5 5
12 0·
0·5
0·5

1
⫺10 5 ⫺10
8

10 9
9
0·5
7

8 5
4 0·
⫺12 7
6
⫺12
0

5
5
3· 0·5
6
1

0·5
⫺14 5
4 3
⫺14
3·5
4 5

0·5

⫺16 3·5 3
⫺16
2
3

3
2·5
2·5
⫺18 2 ⫺18 0·5
0

2
1·5 0·5

5
0

⫺20 1·5
⫺20
0 2 4 6 8 10 0 2 4 6 8 10

Horizontal position relative to pile tip, 2x/B Horizontal position relative to pile tip, 2x/B
(c) (d)

Fig. 17. Post-installation strain distribution, T2-DBS-44: (a) horizontal strain (%); (b)
vertical strain (%); (c) maximum shear strain (%); (d) volumetric strain (%)

Downloaded by [ University Of Western Australia] on [25/05/20]. Copyright © ICE Publishing, all rights reserved.
PLANE-STRAIN MODEL PILE INSTALLATION IN SAND 389

⫺0·5
⫺50
⫺20

⫺1
⫺8
⫺2
1·5
2·5

10
6
2
0
0
2 2
⫺1·5

0 0


4

2
⫺2

3
⫺75
⫺2
⫺1·5

3
⫺2 ⫺2

2 ·5
3 ⫺1·5

15
⫺1

⫺15
⫺6
30
2
0 ⫺0·5 ⫺1


2·5
⫺4 ⫺4

1· 1

3
Vertical position relative to pile tip, 2h/B

Vertical position relative to pile tip, 2h/B


5
0·5

2·5
2

1 1·5
6
10 ⫺0·5
5

0·5 0
⫺6 ⫺6 0

⫺1
1

2
⫺3

⫺0
⫺2·5
⫺8 ⫺8 0·5

·5
5
⫺6
⫺4
0·5
5

4

⫺2
⫺1·5
⫺10 ⫺10 3

1
⫺3 2·5

0·5
1·5
⫺12 ⫺2·5 ⫺12 2

0
⫺2
⫺1

1·5
⫺14 1·
5 ⫺14

0·5
⫺ 1
5
⫺ 0·
1 ⫺ 0 1

0
⫺16 ⫺16
0·5

1
5

0

0
0·5
0·5
⫺18 ⫺18

⫺20 ⫺20
1

0 2 4 6 8 10 0 2 4 6 8 10
Horizontal position relative to pile tip, 2x/B Horizontal position relative to pile tip, 2x/B
(a) (b)
150

1·5
2·5
2·5
2·5
2·5
75
40
25
16

12

0
1
1

2
2 2
9

2·5
1·5

0 0

2·5
12

10
⬎10%
2
20

Section 5.11.3
⫺8
50

⫺2 100
⫺2
2
16

2
75

⫺4
40
25 0

1
3

⫺3
Vertical position relative to pile tip, 2h/B

⫺4
14

⫺4
Vertical position relative to pile tip, 2h/B

0·5
⫺2

1·5
8

⫺1
·5
⫺0·5

⫺6 12
⫺6
9

7
1

20 10 ⫺

5
6

16
0

14 1
0

⫺8 ⫺8
0

12
8

10
9

5
7 1
0·5
0

⫺10 6
⫺10
⫺0·5

8
7 5 4
6 3·5
0

⫺12
1 1

5 4 ⫺12
3

3·5
0·5

4 2·
3 5
⫺14 3·5
2·5 2 ⫺14
0

3
2·5 2
1·5
⫺16
0·5

2 1·5 ⫺16 0

1·5 1
1
⫺18 ⫺18
1

⫺20 ⫺20
0 2 4 6 8 10 0 2 4 6 8 10
Horizontal position relative to pile tip, 2x/B Horizontal position relative to pile tip, 2x/B
(c) (d)

Fig. 18. Post-installation strain distribution, T4-LBS-34: (a) horizontal strain (%); (b)
vertical strain (%); (c) maximum shear strain (%); (d) volumetric strain (%)

in the near field. As these values of strain are referred to the and to the side of the pile in both DBS and LBS. In both
reference frame of the soil element, in zones of high rigid cases the strain level adjacent to the pile is lower than below
body rotation the horizontal and vertical strain axes do not it. If installation is considered as the upward flow of soil
coincide with the x–h axes of the figure. past a stationary pile, then the strain path reversals can be
Bulbs of high horizontal and vertical strain extend below seen as the soil flowing into and out from these bulbs of

Downloaded by [ University Of Western Australia] on [25/05/20]. Copyright © ICE Publishing, all rights reserved.
390 WHITE AND BOLTON
high strain. The upper extent of the bulbs of vertical and 17.63 kN/m3 , indicating a specific volume of 1.53. This is
horizontal strain is bounded by the contour of zero strain. equivalent to a volumetric compressive strain during the test
Above this contour the initial strain probe has been entirely of 32%. The stress-volume path of soil entering the nose
reversed, and the soil is now straining for the first time in cone is sketched on p9–v axes in Fig. 20(c). The mean
the opposite sense to the original probe. pressure, p9, is assumed to equal the measured value of qb .
In contrast to the bulb-shaped distributions of vertical and The soil state lies below the CSL established by Klotz
horizontal strain, the zone of maximum shear strain does not (2000), albeit for a different initial grading of DBS. A
significantly reduce in width as the soil flows past the pile triaxial compression test at a similar mean pressure would
tip. The high shear strain induced as the soil flows to the result in less contraction, with the final state lying on the
558 strain reversal points (Fig. 16) is maintained as the soil CSL. In the case of pile installation, the strain level is
moves past the pile. higher than is achieved during triaxial testing, which could
Comparison of Figs 18 and 19 allows the influence of the lead to additional contraction. Luzzani & Coop (2002) made
ratio of chamber width to pile size to be examined. As there a similar observation from ring shear testing of DBS to a
is good agreement between the shape and the absolute high strain level; a constant volume state was not reached.
values of the contours of strain around each pile, the end The compressive strain on entering the nose cone could
wall boundary effect evidently has negligible influence on be partly balanced by dilation as the soil slips out from the
the kinematic behaviour close to the pile. nose cone, and shears around the shoulder of the pile into
Figures 17 and 18 reveal differing volumetric behaviour in the interface zone, which is at a lower stress level. However,
DBS and LBS. In DBS the zone of compression extends significant particle breakage occurs within the nose cone,
further below the pile. Adjacent to the pile shaft, monotoni- changing the range of attainable voids ratios. The possible
cally decreasing compression with increasing offset is appar- increase in volume as the soil leaves the high-stress zone
ent in DBS. In LBS a zone of net dilation is evident around below the pile tip and undergoes shear strain at low stress
the pile shoulder and close to the pile shaft; a contour of adjacent to the shaft was investigated by measuring the
zero volumetric strain is found at an offset of 2x/B ¼ 1.5, maximum voids ratio of the soil particles recovered by
with compression beyond. This variation of volumetric strain gently shaking the nose cone to break up the interlocked
with offset from the pile shaft is discussed later, after structure. The lowest achievable density was 15.15 kN/m3 ,
observations of the volume change in the interface layer which corresponds to a specific volume of 1.78 (emax ¼
immediately adjacent to the pile have been presented. 0.78). This value is significantly lower than both the maxi-
mum and minimum specific volumes of virgin DBS (Fig.
20(c)). This specific volume represents an upper limit on
COMPARISON WITH THEORETICAL SOLUTIONS dilation of the broken soil. Even if the soil recovered to the
The penetration mechanism characterised by the strain minimum achievable density, a net volume compression of
paths shown in Figs 13 and 14 corresponds qualitatively 21% would remain in the interface zone.
with predictions based on the strain path method (SPM), Full recovery to emax as the soil rounds the pile shoulder
originally proposed by Baligh (1985) and extended by others and forms the interface layer would not occur, as the soil
(Whittle & Baligh, 1988; Teh & Houlsby, 1991; Sagaseta et adjacent to the pile remains under load. An alternative
al., 1997; Gill & Lehane, 2000; Yu et al., 2000). Most stress–volume path would be to assume that the soil moves
published SPM solutions are limited to undrained penetration to a critical state. However, the CSL shown in Fig. 20(c)
in axisymmetry, and so do not capture the volumetric behav- offers no p9–v state at a stress level lower than qb yet at a
iour associated with sand. However, the SPM captures the density attainable by the broken soil; a new CSL must apply.
general trend of a gently increasing vertical and horizontal The actual volumetric strain within the zone of soil adjacent
strain rate, followed by a sharp reversal as soil flows around to the shaft must lie between the value within the nose cone
the pile shoulder. In contrast, cavity expansion methods (32%) and the minimum attainable by the reconstituted soil
predict that each strain quantity will increase monotonically (21%). Therefore the CSL for this heavily overconsolidated
as the pile approaches (or the cavity is expanded), and nose cone sand lies below the CSL for normally consoli-
therefore do not capture the strain history of the soil above dated sand, in p9–v space.
the strain rate reversal lines shown in Fig. 16. Irrecoverable volume reduction was also observed for the
LBS. The soil from the nose cone of test T4-LBS-34 was
sampled in a disturbed state, and found to have a maximum
SOIL COMPRESSION DIRECTLY BELOW PILE specific volume of 1.61. This compares with 1.80 for the
Measurement of the net horizontal displacement ratio for virgin LBS, and represents an irrecoverable volumetric strain
2x0 /B ¼ 1 indicates the volumetric strain within the interface of 10.5%. As with DBS, the actual volumetric strain adja-
zone immediately adjacent to the pile (Fig. 10). The meas- cent to the pile must lie between this value and the greater
ured horizontal displacement ratio of 2˜x/B ¼ 0.8 for 2x0 /B (unknown) strain within the nose cone. These values of
¼ 1 implies a mean volumetric strain of 20% within the compression below the pile tip found from post-test sam-
zone 2xf /B , 1.8, indicating that the interface zone is pling have been added to Figs 17 and 18 to supplement the
significantly more dense than the very near field soil ele- volumetric strains calculated from image analysis.
ments for which volumetric strain paths are shown in Fig. These post-test index measurements indicate that the soil
15. Further evidence of densification close to the pile shaft flowing around the pile tip and into the interface zone
is discussed below. immediately adjacent to the pile shaft suffers significant
A highly compressed region of soil below the pile tip— irrecoverable volumetric compressive strain (.10%), accom-
hereafter referred to as a ‘nose cone’—is visible during all panied by particle breakage. In contrast, Fig. 15 shows that
tests on both DBS and LBS (Fig. 20(a)). Discrete slips of dilation occurs slightly further from the pile tip. These two
soil are observed to slide out from the nose cone and flow modes of volumetric behaviour are compared in Fig. 21, by
around the shaft of the pile. A central core of the nose cone considering the volume changes that are associated with the
is stationary relative to the pile tip, but the shoulders of the stress levels and shear strain encountered close to the pile
zone are not (Fig. 20(b)). tip.
The nose cone from test T3-DBS-71 was sampled during Two streamlines of soil flow are shown in Fig. 21. Soil
the test post-mortem. The density was measured as that flows through the nose cone and forms the interface

Downloaded by [ University Of Western Australia] on [25/05/20]. Copyright © ICE Publishing, all rights reserved.
PLANE-STRAIN MODEL PILE INSTALLATION IN SAND 391

⫺40
⫺20
⫺2
1·5

10
2
0
3

3
⫺ ⫺0·5

2
1 0·
5

1·5
0 0

2
1

⫺1
0·5

1·5

1
⫺0·5

15
1
⫺5 1 ⫺5

0·5 0
2

⫺5

6
10 10 0
⫺8

⫺0·5
⫺1⫺2

5

5
0·5

·5

⫺4
Vertical position relative to pile tip, 2h/B

Vertical position relative to pile tip, 2h/B



⫺10 ⫺10 0·
5

1

⫺1
0
3
5

2·5

0
⫺15 ⫺15 2
⫺2 ·5
⫺1

1·5

⫺20 ⫺
1
⫺20
1

0
·5
⫺0

5

⫺25 ⫺25
5

0·5

0
⫺30 ⫺30
0

0
⫺35 ⫺35
0
0
0
0 0
⫺40 0 ⫺40 0 0 0
0 5 10 15 20 0 5 10 15 20
Horizontal position relative to pile tip, 2x/B Horizontal position relative to pile tip, 2x/B
(a) (b)
50
30
16
12
10

5
4

3
2
8

6
4
4

0 0
5
3

1·5
6
2

6
1·5
8

7
1·5
3

2·5
7
14

1
20

⫺5 ⫺5 0·5
5

1
1

12
0

10 0·
5 0·5
9 0·5
Vertical position relative to pile tip, 2h/B

Vertical position relative to pile tip, 2h/B


3

⫺10 ⫺10
4

8
3·5

7
6

6
5
5

5

⫺15 ⫺15
4 4
5

3

3·5 0
0

3
0

2·5
2

⫺20 ⫺20 0
0

2·5 1·5
0

2
0

2
5

⫺25 ⫺25
1·5
0

1
0

0
1

0·5
⫺30 ⫺30
0

0
0
0

0·5
0
0

0·5 0
0 0
⫺35 ⫺35 0
0

⫺40 ⫺40
0 5 10 15 20 0 5 10 15 20
Horizontal position relative to pile tip, 2x/B Horizontal position relative to pile tip, 2x/B
(c) (d)

Fig. 19. Post-installation strain distribution, T8-DBS-46-narrow: (a) horizontal strain (%); (b)
vertical strain (%); (c) maximum shear strain (%); (d) volumetric strain (%)

layer adjacent to the pile shaft follows streamline ABC. Soil Very high stress and very high shear strain are encoun-
in the very near field, for which PIV data are shown in Figs tered along streamline AB, resulting in volume compression
17–19, follows streamline DEF. A zone of high stress and significant particle breakage, as evident in photographs
corresponding to the photoelastic observations of Allersma of the pile tip (Fig. 20(a)). This volume change is irrecover-
(1988) during plane-strain pile penetration in crushed glass able owing to particle breakage, as demonstrated by the
is shown below the pile tip. index tests described previously. Therefore, as the soil

Downloaded by [ University Of Western Australia] on [25/05/20]. Copyright © ICE Publishing, all rights reserved.
392 WHITE AND BOLTON

Pile

40 mm

Sketched
slip planes

Extent of
`nose cone’

(a) (b)

3·0
Normal compression
line (Coop, 1990)
2·8
Non-linear CSL
(Klotz, 2000)
2·6

2·4
Specific volume, v

Initial state: v ⫽ 2·24


2·2
Maximum specific
volume of soil from
2·0 beneath pile tip after
reconstitution: vmax ⫽ 1·78

1·8

1·6
State of soil below pile tip
v ⫽ 1·53 (measured), qb ⫽ p′: (assumed)
1·4
1 10 100 1000 10000
Mean pressure, p′: kPa
(c)

Fig. 20. Post-mortem sampling of soil below pile tip: (a) ‘nose cone’ of soil beneath pile tip (T4-LBS-34);
(b) slip planes observed within nose cone; (c) stress–volume paths of soil beneath pile tip during test T3-
DBS-71

progresses along the streamline BC, significant volume com-


Density at 2h/B ⬎ 1 pression remains, creating a net increase in density.
Along streamline DE high stress and high shear strain are
Pile
DBS encountered, leading to contraction and some particle break-
Initial density
age. On leaving the zone of high stress, the soil is now
LBS heavily overconsolidated. Also, further shear strain is re-
Offset from pile shaft quired for the soil to pass around the pile shoulder. The
measured strain paths (Figs 13 and 14) show that the shear
C F strain rate remains high within this zone, even though
photoelastic analysis shows that the stress level has dropped.
B Critical state theory predicts that this overconsolidated soil
Low stress will dilate when sheared along streamline EF. Such dilation
E
is seen in Fig. 15. The slight volume compression along DE
is more recoverable than the gross compression and particle
High breakage along AB.
stress The volumetric strain at the end of streamlines ABC
A D
(deduced from index tests) and DEF (deduced from image
analysis) reveals the variation of density with offset from the
Fig. 21. Volumetric behaviour close to pile tip pile shaft. This variation is shown schematically at the top

Downloaded by [ University Of Western Australia] on [25/05/20]. Copyright © ICE Publishing, all rights reserved.
PLANE-STRAIN MODEL PILE INSTALLATION IN SAND 393
of Fig. 21. Adjacent to the pile shaft, the soil has become behaviour was conducted using additional PIV meshes that
more dense, following irrecoverable volume change in the were created in the zone of soil adjacent to the shaft and
nose cone. In DBS the density decreases progressively away beyond the pile tip (h . 0). These meshes were used to
from the pile shaft. In LBS Fig. 18 indicates that this observe the deformation within the soil after passage of the
progressive decrease in density is interrupted by a local pile tip. As the soil close to the pile shaft undergoes only
loose zone in the very near field caused by dilation close to minimal deformation while h . 0, the PIV mesh could
the pile shoulder. extend close to the pile surface. The closest column of PIV
patches was located 4 mm from the pile shaft. Results are
presented from tests T2-DBS-44 and T4-LBS-34.
SOIL FLOW ADJACENT TO PILE SHAFT Figure 22(a) shows the tip of the pile entering the field of
The displacement trajectories shown in Fig. 8 have a view during test T2-DBS-44, with a row of PIV patches
‘tail’, indicating movement of the soil towards the pile shaft established adjacent to the pile shaft. Comparison with a
after the tip has passed. A more detailed investigation of this subsequent image taken after 80 mm (2.5B) of further pene-

h⫽0 h ⫽ 2·5B

(a)

h⫽0

PIV analysis patch

h ⫽ 2·5B

Vector scale: 1 mm

160 140 120 100 80 60 40 20 0


Distance from pile centreline: mm 250 µm
(b)
Horizontal strain during flow from h ⫽ 0 to h ⫽ 2·5B

1·2

1·0

T4-LBS-34
(extension positive, %)

0·8

0·6
T2-DBS-44
0·4

0·2

⫺0·2
160 140 120 100 80 60 40 20 0
Offset from pile centreline: mm
(c)

Fig. 22. Soil movement adjacent to pile shaft: (a) PIV patches established adjacent to pile shaft; (b) measured
displacement vectors; (c) tests T2-DBS-44 and T4-LBS-34, horizontal unloading along a straight pile shaft

Downloaded by [ University Of Western Australia] on [25/05/20]. Copyright © ICE Publishing, all rights reserved.
394 WHITE AND BOLTON
tration allowed the intervening soil movement to be meas- soil interface. The stiff unloading response of the heavily
ured. These vectors of displacement reveal that the soil is overconsolidated soil in the far field is represented by a stiff
moving towards the pile shaft (Fig. 22(b)). The greatest spring (zone A). This spring is fixed in the far field and
vector (250 m, 1.9 pixels) was measured at the patch exerts horizontal stress on the pile shaft. As h increases,
closest to the pile (located 4 mm from the shaft). To elimi- zone B contracts and this spring unloads, reducing the shaft
nate the possibility that this observation was due to a non- friction on the pile.
vertical alignment of the pile, a set of patches on the
opposite side of the pile were tracked. Similar vectors were
obtained. SHAFT FRICTION DEGRADATION: ‘FRICTION
The horizontal strain profiles calculated from the displace- FATIGUE’
ment vectors measured adjacent to the pile during tests T2- It has been widely observed in field and centrifuge model
DBS-44 and T4-LBS-34 are shown in Fig. 22(c). The soil tests that the local shaft friction at a given soil horizon
adjacent to the pile unloads in horizontal extension as the decreases as the pile tip penetrates further. This character-
pile advances and the interface zone contracts. Close-up istic behaviour has been referred to as ‘friction fatigue’
photography during the test post-mortem revealed a 2–3 mm (Heerema, 1980) or the ‘h/R effect’ (Bond & Jardine, 1991),
thick zone of fine broken particles adjacent to the pile shaft. and has been incorporated into recent design methods in an
Figure 23 summarises the observed deformation mechan- empirical manner (Randolph et al., 1994; Jardine & Chow,
ism adjacent to the model pile. This mechanism links the 1996). This behaviour is captured by the mechanism shown
kinematic observation of a contractile interface zone to the in Fig. 24 arising from a contractile interface zone.
degradation of shaft friction close to the pile tip. High Conventional dynamic pile installation methods involve
horizontal stress is created as soil is compressed laterally cyclic shearing at the pile–soil interface rather than the
along streamline XY. As the soil continues along streamline monotonic shearing applied in these tests. The mechanism
YZ the interface zone immediately adjacent to the pile (zone of friction fatigue following interface contraction will be
B in Fig. 24) contracts with continued shearing at the pile– more significant under cyclic loading than under monotonic
loading. It is widely reported from interface shear box
testing that cycles of shear displacement lead to greater
volume contraction than monotonic shearing (Al-Douri &
τs Zone B Poulos, 1991; Tabucanon et al., 1995; Dietz, 2000; DeJong
σ ′h Zone A et al., 2003).
This link between cyclic shear displacement and reduced
Z
shaft friction is reported by Poulos et al. (1988). They note
that the mean shaft friction measured in small-scale model
pile tests after a given number of cycles decreases with
increasing cyclic amplitude. Also, Dolwin et al. (1988)
report a reduction in mean shaft friction with number of
blows (and therefore displacement cycles) required for pile
installation from CAPWAP analysis of driving records from
long offshore piles in calcareous sand.
τmax Zone B

σ ′h,max Zone A
MECHANISM OF PILE SET-UP
Y The mechanism of interface contraction provides the ini-
tial conditions for set-up of displacement piles in sand,
where ‘set-up’ refers to a time-related increase in shaft
capacity, not related to pore pressure change. Immediately
after passing the pile tip, the distribution of radial stress is
as shown by the curve OA in Fig. 24, created as the soil is
pushed outward during flow around the pile tip. As the
interface zone contracts, cylindrical cavity collapse of the
X
stiff overconsolidated soil close to the pile shaft leads to a
sharp reduction in radial stress from the high value created
Fig. 23. Kinematics of friction fatigue close to the pile tip during soil flow around the pile tip. As a result, the radial
stress acting on the pile shaft (point B in Fig. 24) is lower
than beyond the zone influenced by the cavity collapse. Over
A
time, the high gradients in the stress field around the pile
σr relax, creating the radial stress distribution shown as curve
OC, which features an increase in the radial stress acting on
C
the pile shaft and hence set-up.
This mechanism is similar to that proposed by Åstedt
et al. (1992), in which high circumferential stresses, which
Set-up initially ‘arch’ around the pile shaft, relax onto the pile
surface over time, leading to an increase in shaft friction.
B O

Pile CONCLUSIONS
shaft
The results presented in this paper provide detailed experi-
r mental evidence of the penetration mechanism of a displace-
ment pile. A novel image-based deformation measurement
Fig. 24. Radial stress distribution due to interface contraction technique has been used to observe the displacement and

Downloaded by [ University Of Western Australia] on [25/05/20]. Copyright © ICE Publishing, all rights reserved.
PLANE-STRAIN MODEL PILE INSTALLATION IN SAND 395
strain paths induced by the installation of a plane-strain where I is the identity matrix.
model pile into a surcharged calibration chamber. The high The deformation gradient matrix represents the transformation of
precision offered by the measurement technique has allowed a vector from undeformed coordinates, xi , yi to deformed
more detailed analysis of the penetration mechanism than coordinates, x9i , y9i (Fig. 25):
has previously been possible.    
x9i x
A series of eight tests are reported, examining the influ- ¼F i (3)
y9i yi
ence of soil type, initial state, pile breadth and the use of a
driving shoe. The displacement fields and strain paths during Polar decomposition allows the deformation gradient matrix to be
pile installation are found to be relatively independent of divided into strain and rotation components, in order that individual
soil type, although greater compression and a smaller zone components of deformation can be extracted (Chadwick, 1976;
of influence are evident in carbonate sand than in silica Belytschko et al., 2000). The symmetric part of F is the stretch
sand. The shape of the strain path reveals high vertical matrix, U, representing strain, with the skew-symmetric part, R,
compression below the pile tip, followed by horizontal com- containing rotation. (U)2 is known as the right Cauchy-Green strain
pression as the soil element flows around the pile shoulder. matrix. R is antisymmetric, RT R ¼ I, allowing F to be decomposed
This reversal of strain direction is in contrast to cavity as follows:
expansion solutions, which predict monotonic strain paths. 1=2
However, the key features of the penetration mechanism are U ¼ ðF T FÞ (4)
captured by the strain path method. R ¼ FU 1 (5)
The interface zone adjacent to the pile shaft comprised
soil particles that had been broken while passing through the The rotation matrix, R, has the form shown below in equation (6),
zone of high stress below the pile tip, leading to high allowing the rigid body rotation, Ł, during the deformation described
irrecoverable volume reduction. This interface zone was ob- by F to be found. This rotation indicates the orientation of the
material (soil) reference frame, xy, compared with the fixed (camera)
served to contract further while shearing along the pile–soil reference frame, XY.
interface, indicating a mechanism for the degradation of
" #
shaft friction at a given soil horizon with increased pile cos Ł  sin Ł
penetration. This observed contraction has been linked to the R¼ (6)
influence of installation method on pile capacity and the sin Ł cos Ł
possibility of time-related changes in shaft friction (set-up).
The Biot strain matrix, EBiot , is found from the stretch matrix, U,
using the following equation:

APPENDIX 1: STRAIN CALCULATION PROCEDURE EBiot ¼ U  I (7)


In order to calculate strain paths and element rotations the grid of
Biot
PIV patches is divided into a network of triangular elements (Fig. The diagonal elements of E represent the elongation (change in
25). The strain and rotation of the element are calculated as follows. length divided by original length) of line elements oriented with the
The movement of each marker during a single displacement x and y axes—that is, engineering linear strain:
increment is used to construct the displacement gradient matrix, L,
using shape functions (Zienkiewicz, 1967): eng,xy ¼ EBiot
11 (8)
2 3
@u @u eng, yy ¼ EBiot
22 (9)
6 @ X @Y 7
L¼6 4 @v @v 5
7 (1) The sum of the off-diagonal elements is equal to the engineering
shear strain on the xy plane, ª xy :
@ X @Y
The displacement gradient matrix is converted to the deformation ª xy ¼ EBiot Biot
12 þ E 21 (10)
gradient matrix, F, using
2 3 The eigenvalues of EBiot are equal to the principal engineering
@x @x
strains, eng,I , eng,II , with the eigenvectors indicating the inclination
6 @ X @Y 7
F¼6 7 of the principal strain directions to the xy axes.
4 @y @y 5 ¼ L þ I (2)
Natural (logarithmic) strain is defined as the natural logarithm of
@X @Y the current length divided by the original length of a line element.

Node 3
x′3, y′3
u3 , v3
Material (soil element)
reference frame x3, y3

y
Initial position

Displaced position
Y

x Node 2

x′2, y′2
X
u2 , v 2
Node 1
x2, y2
Fixed (camera)
x1, y1
reference frame
u1 , v 1 x′1, y′1

Fig. 25. Triangular element defined by three PIV patches

Downloaded by [ University Of Western Australia] on [25/05/20]. Copyright © ICE Publishing, all rights reserved.
396 WHITE AND BOLTON
The natural strains in the x and y directions are found by taking the Davidson, J. L., Mortensen, R. A. & Barreiro, D. (1981). Deforma-
logarithm of the diagonal elements of U: tions in sand around a cone penetrometer tip. Proc. 10th Int.
log,xx ¼ loge ð U 11 Þ (11) Conf. Soil Mech. Fdn. Engng Stockholm 2, 467–470.
DeJong, J., Randolph, M. F. & White, D. J. (2003). Interface load
log, yy ¼ loge ð U 22 Þ (12) transfer degradation during cyclic loading: a microscale investi-
The engineering principal strain quantities can be converted to gation. Soils Found. 43, No. 4, 81–93.
natural strain using the following equations: Dietz, M. S. (2000). Developing an holistic understanding of inter-
face friction using sand within the direct shear apparatus. PhD
log,I ¼ loge ð1 þ eng,I Þ (13) thesis, University of Bristol.
log,II ¼ loge ð1 þ eng,II Þ (14) Dolwin, J., Khorshid, M. S. & Van Goudoever, P. (1988).
Evaluation of driven pile capacities: methods and results. Proc.
To complete the set of natural strain quantities, maximum natural Int. Conf. Engng for Calcareous Sediments, Perth, Vol. 2, pp.
shear strain is defined as the difference between the natural principal 409–428.
strains: Gill, D. R. & Lehane, B. M. (2000). Extending the strain path
ªlog,max ¼ log,I  log,II (15) analogy for modelling penetrometer installation. Int. J. Numer.
Anal. Methods Geomech. 24, 477–489.
Golightly, C. R. (1989). Engineering properties of carbonate sands.
PhD thesis, Bradford University.
REFERENCES Golightly, C. R. & Hyde, A. F. L. (1988). Some fundamental
Airey, D. W., Randolph, M. F. & Hyden, A. M. (1988). The strength properties of carbonate soils. Proc. Int. Conf. Calcareous Sedi-
and stiffness of two calcareous sands. Proc. Int. Conf. on Engng ments, Vol. 1, pp. 69–78.
for Calcareous Sediments, Perth, Vol. 1, pp. 43–50. Heerema, E. P. (1980). Predicting pile driveability: heather as an
Al-Douri, R. H. & Poulos, H. G. (1991). Static and direct shear illustration of the friction fatigue theory. Ground Engng 13, 15–37.
tests on carbonate sands. ASTM Geotech. Test. J. 15, No. 2, Houlsby, G. T., Evans, K. M. & Sweeney, M. A. (1988). End
138–157. bearing capacity of model piles in layered carbonate soils. Proc.
Allersma, H. G. B. (1988). Photoelastic investigation of the stress Int. Conf. on Engng for Calcareous Sediments, Perth, Vol. 1, pp.
distribution during penetration. Proc. 2nd Eur. Symp. Penetration 209–214.
Testing, Amsterdam, 411–418. Houlsby, G. T. & Hitchman, R. (1988). Calibration chamber
API (2000). Recommended practice of planning, designing and tests of a cone penetrometer in sand. Géotechnique 38, No. 1,
constructing fixed offshore platforms: working stress design, 39–44.
RP2A, 20th edn, pp. 59–61. Washington: American Petroleum Jardine, R. J. & Chow, F. C. (1996). New design methods for
Institute. offshore piles, MTD Publication 96/103. London: Marine Tech-
Åstedt, B., Weiner, L. & Holm, G. (1992). Increase in bearing nology Directorate.
capacity with time for friction piles in silt and sand. Proceed- Jardine, R. J., Symes, M. J. & Burland, J. B. (1984). The measure-
ings of Nordic Geotechnical Meeting, pp. 411–416. ment of soil stiffness in the triaxial apparatus. Géotechnique 34,
Atkinson, J. H., Richardson, D. & Stallebrass, S. E. (1990). Effect No. 3, 323–340.
of recent stress history on the stiffness of overconsolidated soil. Jardine, R. J., Standing, J. R., Jardine, F. M., Bond, A. J. & Parker,
Géotechnique 40, No. 4, 531–540. E. (2001). Competition to assess the reliability of pile prediction
Baligh, M. M. (1985). Strain path method. ASCE J. Geotech. Engng methods. Proc. 15th Int. Conf. Soil Mech. Geotech. Engng,
111, No. 9, 1108–1136. Istanbul 2, 911–914.
Belytchko, E., Liu, W. K. & Moran, B. (2000). Non-linear finite Jovičič, V. & Coop, M. R. (1997). Stiffness of coarse-grained soils
elements for continua and structures. Chichester: Wiley. at small strains. Géotechnique 47, No. 3, 545–561.
Berezantzev, V. C., Kristoforov, V. & Golubkov, V. (1961). Load- Klotz, E. U. (2000). Influence of state on the capacity of driven
bearing capacity and deformation of piled foundations. Proc. 5th piles in sand. PhD thesis, City University, London.
Int. Conf. Soil Mech. Fdn. Engng, Paris 2, 11–12. Klotz, E. U. & Coop, M. R. (2001). An investigation of the effect
Bolton, M. D., Gui, M. W., Garnier, J., Corte, J. F., Bagge, G., of soil state on the capacity of driven piles in sands. Géotechni-
Laue, J. & Renzi, R. (1999). Centrifuge cone penetration tests que 51, No. 9, 733–751.
in sand. Géotechnique 49, No. 4, 543–552. Kraft, L. M. (1990). Computing axial pile capacity in sands for
Bond, A. J. & Jardine, R. J. (1991). Effects of installing displace- offshore conditions. Marine Geotech. 9, 61–72.
ment piles in a high OCR clay. Géotechnique 41, No. 3, 341– Lee, J. H. & Salgado, R. (1999). Determination of pile base
363. resistance in sands. ASCE J. Geotech. Geoenviron. Engng 125,
Borghi, X., White, D. J., Bolton, M. D. & Springman, S. M. (2001). No. 8, 673–683.
Empirical pile design based on CPT results: an explanation for Lee, S. Y. (1989). Centrifuge modelling of cone penetration testing
the reduction of unit base resistance between CPTs and piles. in cohesionless soils. PhD thesis, Cambridge University.
Proc. 5th Int. Conf. on Deep Foundation Practice, Singapore, Lunne, T., Robertson, P. K. & Powell, J. J. M. (1997). Cone
125–132. penetration testing in geotechnical practice. London: Blackie.
Bowman, E. T., Soga, K. & Drummond, W. (2001). Particle shape Luzzani, L. & Coop, M. R. (2002). On the relationship between
characterisation using Fourier descriptor analysis. Géotechnique particle breakage and the critical state of sands. Soils Found. 42,
51, No. 6, 545–554. No. 2, 71–82.
BSI (1990). British Standard methods of tests for soils for civil Mikasa, M. & Takada, N. (1973). Significance of centrifugal model
engineering purposes, BS 1377. Milton Keynes: British Stan- tests in soil mechanics. Proc. 8th Int. Conf. Soil Mech. Fdn.
dards Institution. Engng, Moscow, 273–278.
Bruno, D. (1999). Dynamic and static load testing of driven piles in Poulos, H. G., Randolph, M. F. & Semple, R. M. (1988). Evaluation
sand. PhD thesis, University of Western Australia. of pile friction from conductor tests. Proc. Int. Conf. Calcareous
Budhu, M. (1979). Simple shear deformation of sands. PhD thesis, Sediments, Perth, Vol. 2, pp. 599–605.
University of Cambridge. Randolph, M. F., Dolwin, J. & Beck, R. (1994). Design of driven
Chadwick, P. (1976). Continuum mechanics: concise theory and piles in sand. Géotechnique 44, No. 3, 427–448.
problems. London: George Allen & Unwin. Robinsky, E. I. & Morrison, C. F. (1964). Sand displacement and
Chong, F. (1988). Density changes of sand on cone penetration compaction around model friction piles. Can. Geotech. J. 1, No.
resistance. Proc. 1st Symp. Penetration Testing, ISOPT-1, 2, 2, 81–93.
707–714. Rotterdam: Balkema. Sagaseta, C., Whittle, A. J. & Santagata, M. (1997). Deformation
Chow, F. C. (1997). Investigations into the behaviour of displace- analysis of shallow penetration in clay. Int. J. Numer. Anal.
ment piles for offshore foundations. PhD thesis, University of Methods Geomech. 21, No. 10, 687–719.
London (Imperial College). Salgado, R., Mitchell, J. K. & Jamiolkowski, M. (1997). Cavity
Coop, M. R. (1990). The mechanics of uncemented carbonate expansion and penetration resistance in sand. ASCE J. Geotech.
sands. Géotechnique 40, No. 4, 607–626. Geoenviron. Engng 123, No. 4, 344–354.

Downloaded by [ University Of Western Australia] on [25/05/20]. Copyright © ICE Publishing, all rights reserved.
PLANE-STRAIN MODEL PILE INSTALLATION IN SAND 397
Schnaid, F. (1990). A study of the cone-pressuremeter test in sand. White, D. J., Take, W. A. & Bolton, M. D. (2003). Soil deformation
DPhil thesis, Oxford University. measurement using particle image velocimetry (PIV) and photo-
Sentenac, P., Lynch, R. J. & Bolton, M. D. (2001). Measurement of gammetry. Géotechnique 53, No. 7, 619–631.
the side-wall boundary effect in soil columns using fibre-optics White, D. J. & Bolton, M. D. (2004). A review of field measure-
sensing. Int. J. Physical Modelling in Geotechnics 1, No. 4, 35–41. ments of CPT and pile base resistance. Proc. Instn Civ. Engrs,
Stroud, M. A. (1971). Sand at low stress levels in the simple shear Geotech. Engng, forthcoming.
apparatus. PhD thesis, Cambridge University. Whittle, A. J. & Baligh, M. M. (1988). The behaviour of piles
Tabucanon, J. T., Airey, D. W. & Poulos, H. G. (1995). Pile skin supporting tension leg platforms, Final MIT Sea Grant & Joint
friction in sands from constant normal stiffness tests. ASTM Industry Consortium Report. Department of Civil Engineering,
Geotech. Test. J. 18, No. 3, 350–364. Massachusetts Institute of Technology.
Tan, F. S. C. (1990). Centrifuge and theoretical modelling of Winterkorn, A. F. & Fang, S. Y. (1975). Foundation engineering
conical footings on sand. PhD thesis, Cambridge University. handbook. New York: Van Nostrand Reinhold.
Teh, C. I. & Houlsby, G. T. (1991). Analytical study of the cone Yasufuku, N. & Hyde, A. F. L. (1995). Pile end-bearing capacity in
penetration test in clay. Géotechnique 41, No. 1, 17–34. crushable sands. Géotechnique 45, No. 4, 663–676.
Tejchman, A. & Gwizdala, K. (1979). Analysis of safety factors of Yu, H. S., Herrman, L. R. & Boulanger, R. W. (2000). Analysis of
bearing capacity for large diameter piles. Proc. Eur. Conf. Soil steady cone penetration in clay. ASCE J. Geotech. Geoenviron.
Mech. Fdn Engng, Brighton 1, 293–296. Engng 126, No. 7, 594–605.
White, D. J. (2002). An investigation into the behaviour of pressed- Zienkiewicz, O. C. (1967). The finite element method in structural
in piles. PhD thesis, University of Cambridge. and continuum mechanics. London: McGraw-Hill.

Downloaded by [ University Of Western Australia] on [25/05/20]. Copyright © ICE Publishing, all rights reserved.

You might also like