You are on page 1of 11

DYNAMIC AND STATIC LOAD TESTING OF MODEL PILES DRIVEN INTO

DENSE SAND
By D. Bruno1 and M. F. Randolph2

ABSTRACT: Systematic studies of pile response during static and dynamic load tests are generally too expensive
to conduct in the field, and at model scale may be limited by scaling effects and the ability to obtain accurate
stress waves. This paper describes such a study, conducted at model scale in a geotechnical centrifuge, for piles
driven into dense sand. Adverse scaling effects were minimized by the use of extremely fine sand (silica flour),
and accurate stress waves were obtained using high-frequency data logging, together with a Hopkinson bar
arrangement for the measurement of pile-head velocity. The overall aim of the study was to compare dynamic
and static test data, for open- and closed-ended piles driven into dense sand, for a range of delivered hammer
energies. Open- and close-ended model piles were driven into dense sand and statically load tested at different
Downloaded from ascelibrary.org by Aalborg University on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

penetrations, without stopping the centrifuge. Stress-wave data were collected, during continuous driving and
from single blows immediately prior to the commencement of static load testing. An assessment of the accuracy
of the mobilized soil resistance estimated from dynamic testing, using different levels of analysis, has been made
by direct comparison with the static load test data. Particular emphasis has been placed on the performance of
the dynamic analyses in light of different driving conditions and delivered hammer energy. From measurements
of the delivered hammer energy, the efficiency of the centrifuge pile driving system was also assessed.

INTRODUCTION tests arises for open-ended piles, where the mode of penetra-
The assessment of the static bearing capacity of piles using tion under dynamic conditions, with soil moving up inside the
dynamic testing is of major interest because of the lower cost open pipe pile, differs from that during static loading, where
of dynamic testing, and also in situations such as offshore the pile will generally behave as fully plugged. It might be
operations, where static loading is generally not feasible. Com- expected that dynamic testing would significantly underesti-
parisons between dynamic and static estimates of pile capacity mate the fully plugged pile capacity under static loading, par-
have been reported widely in the literature (Broms 1981, 1985; ticularly under the relatively high-energy levels needed to
Fellenius 1988; Barends 1992; Goble and Likins 1996; Town- cause sufficient displacement of the pile under dynamic con-
send 1996). Open-ended piles are an area of particular diffi- ditions.
culty because of the different modes of penetration under This paper presents results from an experimental study of
dynamic and static loading conditions. Systematic studies are pile capacity under dynamic and static conditions, with the
needed in uniform soil conditions, where differences between main focus being the effect of delivered hammer energy on
dynamic and static loading can be evaluated for both open- the dynamic response (and estimated soil resistance), and the
and closed-ended piles, for a range of delivered hammer effect of the pile tip conditions (open- and closed-ended) on
energy. the dynamic and static responses. Both compression and ten-
The high cost of field testing necessitates that such studies sion load tests were performed, at varying pile embedment
are undertaken, at least initially, at model scale. An extensive ratios. However, the present paper is restricted to a comparison
centrifuge testing program has been performed with the aim of the responses in compression only.
of examining the drivability of open- and closed-ended piles
in dense sand, and comparing capacities measured under both CENTRIFUGE MODELING
dynamic and static loading conditions. Because it is difficult, Centrifuge modeling has a number of advantages over al-
at model scale, to obtain accurate measurements of the high ternative modeling techniques in geotechnical engineering, al-
accelerations associated with pile driving, the pile velocity lowing a wide range of soil and pile parameters to be varied
during a hammer blow has been determined using a Hopkin- independently while maintaining correct scaling of the self-
son bar arrangement (Zelikson 1991; Bruno and Randolph weight stress gradient in the soil inside and outside the model
1998). In this approach, the pile velocity is deduced directly pile. Correct reproduction of internal and external stresses ad-
from the stress wave traveling in a light plastic rod attached jacent to the pile wall allows an accurate assessment of the
near the pile head. Measurement of both force and velocity at mobilization and degradation of shaft and end-bearing resis-
the pile head during hammer blows allows dynamic analyses tance during pile installation and loading.
of varying degrees of sophistication to be undertaken to esti- The experimental work presented in this paper was carried
mate the static capacity of the model piles, in addition to as- out on the fixed beam geotechnical centrifuge facility at the
sessing impact energy and hammer efficiency. The deduced University of Western Australia. The facility houses an Acu-
pile capacities may then be compared with results from direct tronic 661 centrifuge with a capacity of 40 g-tonnes and a
static loading of the pile. platform radius of 1.8 m. A detailed account of the geotech-
A potential problem in comparing dynamic and static load nical centrifuge and associated equipment may be found in
1
PhD Student, Sp. Res. Ctr. for Offshore Found. Sys., Univ. of Western
Randolph et al. (1991). The tests conducted here were per-
Australia, Nedlands 6907, Perth, Australia. formed in a strongbox of width 390 mm, length 650 mm, and
2
Prof. of Civ. Engrg., Sp. Res. Ctr. for Offshore Found. Sys., Univ. of height 325 mm, containing between 260 and 275 mm depth
Western Australia, Nedlands 6907, Perth, Australia. of soil.
Note. Discussion open until April 1, 2000. To extend the closing date Centrifuge modeling laws have been discussed in general
one month, a written request must be filed with the ASCE Manager of terms by Schofield (1980), and in the context of modeling pile
Journals. The manuscript for this paper was submitted for review and
possible publication on October 9, 1998. This paper is part of the Journal
driving by de Nicola (1996). Relevant scaling relationships are
of Geotechnical and Geoenvironmental Engineering, Vol. 125, No. 11, summarized in Table 1. It may be seen that all accelerations
November, 1999. 䉷ASCE, ISSN 1090-0241/99/0011-0988–0998/$8.00 are scaled by the modeling ratio N. The tests described here
⫹ $.50 per page. Paper No. 19420. were undertaken at a scale of 1:100, so that acceleration levels
988 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING / NOVEMBER 1999

J. Geotech. Geoenviron. Eng., 1999, 125(11): 988-998


TABLE 1. Scaling Ratios for Centrifuge Modeling
Quantity Scaling ratio
(1) (2)
Stress 1
Strain 1
Acceleration N
Length 1/N
Velocity 1
Force 1/N2
Time (inertia) 1/N
Time (diffusion) 1/N2

during impact are approximately 100 times higher than at field


scale, and may reach levels of 15,000–40,000g. The disparity
between scaling of time for inertia and diffusion events (such
Downloaded from ascelibrary.org by Aalborg University on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

as consolidation) was partly compensated for in the present


tests by the use of very fine-grained soil, with much lower
permeability than a typical field sand.
FIG. 2. CPT Data (All Test Samples)

SOIL SAMPLE TABLE 3. Centrifuge Sample Properties


Model testing allows preparation of reasonably uniform soil Property Test C Test E Test H Test I Test J
samples under controlled conditions of density, water table, (1) (2) (3) (4) (5) (6)
etc. However, there is a potential problem with scale effects Bulk unit weight, ␥bulk 20.0 19.4 19.7 19.8 19.2
arising from the relative magnitudes of the soil grains and the Water content, w(%) 23 24 24 23 25
primary structural dimensions of the model piles, such as di- Voids ratio, e 0.60 0.67 0.64 0.63 0.70
ameter and wall thickness. This problem has been minimized Relative density, ID (%) 95 86 90 91 82
by reducing the grain size of the sand medium to a fine-grained
silica flour or silt (de Nicola 1996), with a mean particle size
d50 of 45 ␮m and a d10 of 6 ␮m. The well-graded nature of A comprehensive series of constant normal stiffness direct
the soil, as shown by the particle size distribution curve in Fig. shear tests were conducted by Maslen (1997) to assess the
1, allows it to be densely compacted, with the potential to gain frictional and dilatational characteristics of the silica flour used
very high strengths. All centrifuge tests reported here were in the centrifuge tests. Similar tests were conducted on coarser
conducted in the silica flour, under normally consolidated, sat- (field scale) sand and it was concluded that, while the silica
urated conditions. Table 2 summarizes key properties of the flour had a mean particle size an order of magnitude smaller
soil. than the coarser sand, it correctly modeled the mechanical
properties of a field scale sand.
Within the centrifuge, the variation of cone resistance qc
with depth was determined using a 7-mm-diameter model cone
penetrometer (CPT). Up to four CPTs were conducted in any
individual soil sample. There was excellent reproducibility
between the CPTs performed in each sample, indicating a
high level of (lateral) homogeneity. Fig. 2 shows the average
CPT profiles for the five different soil samples that were used
for the dynamic and static pile load tests presented in this
paper. Generally, the trend is for qc to increase with depth until
a prototype depth of around 17 m, where the proximity of
the strongbox base (at about 23 m) led to increasing gradients
of qc.
Because the silica flour is too fine for pluviation, each sam-
ple was prepared by pouring a saturated slurry into the strong-
box, and then densifying the sample on a vibrating table to
achieve a final (average) density. Some variations in soil den-
sity, due to different durations of vibration, are evident from
the cone profile. It is also evident that the vibration table
FIG. 1. Particle Size Distribution Curve for Silica Flour tends to concentrate the densification in the bottom third of
the sample. By placing a perforated board with a 20-kg mass
TABLE 2. Mechanical Properties of Silica Flour on it (Sample C), a more uniformly densified sample was
Property Value produced. For Sample I, no vibration was used, and as a
(1) (2) result, a degree of nonhomogeneity was present in the central
band of the sample. Table 3 summarizes the average density
Mean particle size, d50 45 ␮m
Void ratio, e 0.57–1.30
of each sample.
Minimum saturated density, ␥min 16.9 kN/m3
Maximum saturated density, ␥max 22.0 kN/m3 PILE TESTING APPARATUS
Permeability, k 2 ⫻ 10⫺6 m/s
Specific gravity, Gs 2.66 Miniature Pile Driving Hammer
Critical friction angle, ␾ ⬘cv 37.8 degrees
Steel/soil friction angle, ␦s 29.2 degrees
In the centrifuge, the model piles have been installed in-
flight using a pneumatically operated drop hammer, which
JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING / NOVEMBER 1999 / 989

J. Geotech. Geoenviron. Eng., 1999, 125(11): 988-998


simulates a prototype drop hammer of 28 tonnes (ram mass),
such as would be used for an equivalent prototype pile.
The hammer may operate continuously, or with single blows
to permit stress-wave data measurements. de Nicola and Ran-
dolph (1994) have given a more detailed description of the
miniature pile driving actuator.

Model Piles
A detailed description of the model piles has been given by
Bruno (1998), so only the key features are summarized here.
Fig. 3 shows a schematic diagram of the model piles used
throughout the centrifuge testing program. At 100g the model
piles have a prototype diameter of 0.95 m, with a wall thick-
ness of 0.05 m and a maximum embedment depth of 20 m.
Downloaded from ascelibrary.org by Aalborg University on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Both piles are fabricated from a stainless-steel section that has


been sandblasted to produce a lightly roughened surface. The
pile roughness R was measured with a diamond stylus to pro-
duce a centerline average of 2.2 ␮m.
For the model pile roughness to correctly scale with the
prototype roughness, the ratio of the mean sand particle size FIG. 4. Interface Friction Angles and Normalized Roughness
d50 and the steel roughness was conserved, with a normalized (after Kishida and Uesugi 1987; Jardine et al. 1992)
roughness R/d50 of 48.8 ⫻ 10⫺3. The normalized roughness of
the model pile, and corresponding peak and residual soil-pile pile and through a series of pulleys to a potentiometer (de
interface angles, are compared in Fig. 4 with studies performed Nicola 1996). Piles were also installed closed-ended, sealing
by Kishida and Uesugi (1987) and Jardine et al. (1992). The the toe of the piles with a flat end-cap that was flush with
good agreement between the silica flour, used in the centrifuge respect to the outside diameter of the pile.
tests, and the (often larger) soils tested by the two studies
indicates that an appropriate representation of the frictional
Hopkinson Bar Arrangement
interaction at the pile/soil interface was achieved in the model.
The uninstrumented pile (UP) has only one set of strain For model pile tests in the centrifuge, the transient force of
gauges, near the pile head, to provide force measurements dur- a stress wave may be measured with strain gauges. However,
ing dynamic pile testing as well as total bearing capacity mea- it is difficult to miniaturize the accelerometer needed for ve-
surements during static loading. On the other hand, the instru- locity measurements, owing to the high levels of acceleration
mented pile (IP) has seven levels of axial strain gauges, which (requiring accelerometers capable of withstanding 100,000g)
provided additional information regarding the distribution of and distortion of any measurements due to the mass of the
shaft friction along the embedded length of the pile during instruments themselves. This problem was overcome, using a
static load testing. These gauges were glued to the outside technique known as Hopkinson bar (Zelikson 1991).
of the pile, protected by a uniform, but sandblasted, layer of The Hopkinson bar itself is a thin rectangular strip of PVC
epoxy. that is attached adjacent to the strain gauges, at the pile head.
For piles installed open-ended, progression of the soil plug To measure strain, the PVC strip is strain gauged with two
was monitored by a lightweight mass resting on the soil plug diametrically opposed, high-resistance strain gauges located at
and attached to a kevlar line that passed up the inside of the the end attached to the model pile. A schematic diagram of
the arrangement is shown in Fig. 5.
As the point of attachment moves, a wave is generated in
the strip, which travels up to the end of the strip where it is
reflected. By virtue of the low wave speed in the PVC, no
reflected waves reach the strain-gauged end of the strip until
the main stress-wave response within the pile is complete. This

FIG. 3. Schematic of Model Piles FIG. 5. Schematic of Hopkinson Bar Arrangement

990 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING / NOVEMBER 1999

J. Geotech. Geoenviron. Eng., 1999, 125(11): 988-998


allows measurement of a one-directional stress wave through with fixed prototype drop heights ranging between 1.0 and 1.7
the strip during the impact of the ram mass on the pile. The m. Once the desired penetration depth was achieved, a series
one-directional wave induced in the PVC is a crucial factor of dynamic load tests were followed by static tests.
because the pile-head velocity v is only proportional to strain The dynamic tests were performed several minutes after
for unidirectional waves (upward or downward, not both). The driving, by striking the piles with single blows from the ram
pile-head velocity may be calculated from the following re- mass. Often, at any one penetration, individual blows were
lationship: delivered and recorded from different drop heights.
For the UP, static tests were performed immediately after
v = εc (1) the dynamic tests, generally in compression first, followed by
tension, although this order was reversed for some tests. To
where ε = strain; and c = wave speed in the PVC.
assess the mobilization of local shaft friction and end bearing,
Because the wave speed in the 200-mm-long steel model
where possible, the IP was installed to the same penetration
pile is around 5,000 m/s, a high-speed data acquisition system
depths at a neighboring site and tested statically. Few dynamic
is required to capture the dynamic event, which has a return
tests were conducted on the IP because early testing indicated
time of 90 ␮s. This is achieved using a card manufactured by
that the presence of the epoxy coating corrupted the quality
Downloaded from ascelibrary.org by Aalborg University on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Bakker, in a standard personal computer. The card has a max-


of the dynamic data (presumed to be due to the different wave
imum allowable sampling rate for the two channels (pile gauge
speeds in epoxy and steel).
and the PVC strip gauge) of 400 kHz, so that the data may be
discretized at intervals of 2.5 ␮s.
PILE TEST DATA
PILE TESTING PROCEDURE
Dynamic and static tests were performed on 36 piles, 22
Prior to any pile installation and testing, two CPTs were open-ended and 14 closed-ended, at prototype penetration
performed in each sample at a minimum of 10 pile diameters depths varying between 4.7 and 20 m. Each test indicator is
D from the nearest boundary or pile test site. Up to seven piles represented by a letter followed by two numbers, signifying
could be installed along the centerline of the testing container the test sample (C, E, H, I, and J), test drive (between 1 and
while ensuring a minimum separation of 7D between sites. The 6 in any sample), and test depth (between 1 and 3 at any site).
piles were driven continuously with the miniature pile driving A summary of the pile tests is presented in Table 4. All pre-
actuator, which delivered the 28-tonne (prototype) ram mass sented data are discussed in terms of prototype units.

TABLE 4. Pile Test Details


Test Dynamic test Plug Permanent
Test depth Pile toe drop heights height set Qstatic
indicator (m) condition Static tests (m) (m) (mm) (MN)
(1) (2) (3) (4) (5) (6) (7) (8)
C1-1 7.60 Open C-T-C 1 5.14 7.84 7.27
C1-2 14.00 Open C-T-C 1.7 9.20 3.99 12.91
C1-3 18.30 Open C-T-C 1.7 10.50 2.80 10.50
C2-1 6.00 Closed C-T-C 1 — 7.80 8.88
C4-1 7.44 Open C-T-C 1 5.68 3.33 7.12
C4-2 14.00 Open C-T-C 1.7 8.60 4.44 12.32
C4-3 18.30 Open C-T-C 1.7 9.70 1.31 11.20
E6-1 6.61 Open C-T-C 1.7 2.76 3.88 7.61
E6-2 12.56 Open T-C-T 1.7 6.40 1.97 10.07
H2-1 4.86 Open C-T-C-T 0.9 2.20 2.88 5.68
H2-2 12.33 Open C-T-C-T 1, 1.2 12.33 4.03 8.10
H2-3 17.33 Open C-T-C-T 1, 1.2, 1.5, 1.7 13.70 4.26 11.49
H3-1 4.71 Open T-C-T-C 1.6 1.92 2.86 5.11
H3-2 13.75 Open T-C-T-C 1, 1.2, 1.5, 1.7 10.31 7.87 9.28
H3-3 18.76 Open T-C-T-C 1, 1.2, 1.5, 1.7 15.36 5.10 12.50
H5-1 4.86 Closed C-T-C-T 1, 1.2, 1.7 — 5.81 7.43
H5-2 12.57 Closed C-T-C-T 1.2 — 8.20 8.00
H6-1 4.67 Closed T-C-T-C 1, 1.2, 1.5, 1.7 — 3.33 5.60
I2-1 7.91 Open C-T-C-T 0.9, 1.1, 1.4, 1.6 3.40 3.22 9.62
I2-2 12.13 Open C-T-C-T 0.6, 0.8, 1, 1.3, 1.5, 1.6 7.01 0.31 13.52
I3-1 9.08 Closed C-T-C-T 0.8, 1, 1.2, 1.5, 1.7 — 1.80 7.92
I3-2 13.38 Closed C-T-C-T 0.8, 1, 1.2, 1.5, 1.7 — 0.46 11.52
I3-3 16.78 Closed C-T-C-T 0.8, 1, 1.2, 1.5, 1.7, 1.8 — 0.44 17.01
I4-1 7.76 Closed C-T-C-T 0.8, 1, 1.2, 1.5, 1.7, 1.8 — 2.99 10.79
I4-2 11.96 Closed C-T-C-T 0.8, 1, 1.2, 1.7, 1.8 — 1.84 13.99
I4-3 17.37 Closed C-T-C-T 0.8, 1, 1.2, 1.5, 1.7, 1.8 — 0.31 16.95
J1-1 10.73 Open C-T-C-T 1, 1.2, 1.5, 1.7, 1.9 6.30 2.81 6.20
J1-2 14.93 Open C-T-C-T 1, 1.2, 1.5, 1.7, 1.8 10.00 2.22 10.12
J1-3 20.03 Open C-T 1, 1.2, 1.5, 1.7, 1.8 14.40 0.60 16.37
J2-1 7.15 Open C-T-C-T 0.8, 1.1, 1.4, 1.6, 1.8 4.10 4.37 5.74
J2-2 12.05 Open C-T-C-T 1, 1.2, 1.5, 1.7 8.27 3.06 9.40
J2-3 18.60 Open C-T-C-T 1, 1.2, 1.5, 1.7 12.94 0.67 13.38
J3-1 10.58 Closed C-T-C-T 1, 1.2, 1.5, 1.7, 1.8 — 2.25 9.09
J3-2 14.79 Closed C-T-C-T 1, 1.2, 1.5, 1.7, 1.8 — 0.31 12.15
J4-1 7.17 Closed C-T-C-T 1, 1.2, 1.5, 1.7 — 2.23 5.12
J4-2 12.04 Closed C-T-C-T 0.8, 1, 1.2, 1.5, 1.7 — 1.80 8.48
Note: C = compression; T = tension; Qs = taken at pile-head displacement of 10% of pile diameter.

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING / NOVEMBER 1999 / 991

J. Geotech. Geoenviron. Eng., 1999, 125(11): 988-998


Pile Installation As expected, considerably more blows are required to install
the piles for a 1-m drop height as opposed to a 1.7-m drop
As described earlier, all model piles were driven into the height. For a 1.7-m drop height, the closed-ended pile required
samples with the miniature pile driving hammer system using slightly more blows than the open-ended piles. Typically, the
a 28-tonne prototype drop hammer. A preset drop height was tests have shown that close-ended piles require around 25%
maintained throughout driving, usually starting with a 1-m more blows than open-ended piles, on average.
drop height for the first 5 m and then increasing to 1.7 m until
ultimate pile penetration. Soil Plug Formation
Several driving records from the installation of the model
piles are shown in Fig. 6. In Fig. 6(a), the soil strength stra- Generally, all open-ended piles remained partially plugged
tigraphy (reflected in the CPT data) follows a similar profile throughout installation, with Fig. 7 showing the typical de-
to the driving record of the open-ended pile. Increasing drop velopment of the internal soil plug with depth. Due to a high
heights were necessary as the pile became harder to install. level of reproducibility in each sample, only a single curve has
Fig. 6(b) shows the cumulative hammer blows measured dur- been shown to represent the average of several measured plug
ing driving for three open- and one closed-ended pile driven lengths. A linear fit has been added to the data to emphasize
Downloaded from ascelibrary.org by Aalborg University on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

into Sample E. Installation effects due to different hammer the distinct transition in the incremental filling ratio (IFR),
drop heights and pile toe conditions were examined in this after a depth of around 5 m, where the IFR is given by (Brucy
sample. At an embedment depth of around 5.5 m, the drop et al. 1991)
height for two of the open-ended piles was increased to 1.7
m, whereas 1 m was maintained for the third open-ended pile. change in plug height
IFR = (2)
change in pile penetration
This is believed to be due primarily to the change in drop
height, from 1 to 1.7 m, at which the ram mass is delivered
during driving. Brucy et al. (1991) also observed that the IFR
was strongly affected by the input energy, with increased en-
ergy leading to a higher IFR.
It is interesting to observe that, over most of the installation
with a 1.7-m drop height, the IFR is steady at about 0.9. Per-
haps the main factor that limits the IFR to 0.9 is the relatively
large pile wall thickness (or small internal pile diameter) com-
pared with the mean soil grain size. This may be responsible
for inhibiting the initial ingression of soil into the pile.

Static Pile Tests


For simplicity, the bearing capacity Qstatic was taken as the
load required to displace the pile head by 10% of the pile
diameter (or 0.1D). Applying the Davisson criterion, where the
displacement required for Qstatic is 0.1D ⫹ PL/EA [see (4)],
increases the quoted static pile capacities by <8%. Fig. 8
shows typical load-displacement data measured from static
tests performed on an open-ended pile in Sample J at three
different penetrations at the same site. The static tests were
performed in compression followed by tension, and then cy-
cled twice more. In compression, there is a clear tendency for
the bearing capacity to continue to increase with increasing
pile displacement. Conversely, in tension, the peak tensile load

FIG. 6. (a) Drivability of Open-Ended Pile (Test I); (b) Driving


Record for Open- and Closed-Ended Piles (Test E) FIG. 7. Formation of Soil Plug during Driving, All Tests

992 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING / NOVEMBER 1999

J. Geotech. Geoenviron. Eng., 1999, 125(11): 988-998


pile was installed in a partially plugged mode. This transition
from plugged to partially plugged behavior has also been ob-
served at model scale (O’Neill and Raines 1991; Paik and Lee
1993; de Nicola 1996), as well as at field scale (Brucy et al.
1991).

Dynamic Pile Tests


Typical data measured from dynamic tests conducted on
open- and closed-ended piles, embedded at a depth of 7.15 m,
are shown in Figs. 9 and 10, respectively. The large mismatch
between the force and Zv curves (product of pile impedance
Z = EA/c and velocity v, where E is Young’s modulus, A is
the net cross-sectional area of the pile, and c the wave speed),
is due to the large change in impedance that is present 30 mm
(prototype return time of 1.2 ms) from the point of instru-
Downloaded from ascelibrary.org by Aalborg University on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

mentation. The impedance of the model pile is 595 Ns/m com-


pared with only 16 Ns/m for the Hopkinson bar (Bruno and
Randolph 1998). Therefore, any distortions in the proportion-
ality between the force and Zv curves due to the Hopkinson
FIG. 8. Typical Static Test Data (Drive J 2, Open-Ended Pile) bar itself would be minimal because the combined impedance
of the pile and the Hopkinson bar is only 2.7% greater than
is usually achieved within a displacement of 0.1D with the that of the pile alone.
displacement required to mobilize the peak tensile load in- The extent of the stress-wave reflection from the pile toe,
creasing with pile penetration. at a return time of 2L/c, is normally characterized by the ten-
The majority of the soil resistance mobilized by the model sile dip in force near the pile head and a corresponding local
piles was end-bearing resistance, comprising 75, 65, and 60% maximum in pile-head velocity. The magnitude of these per-
of the total bearing capacity of the piles for tests conducted at turbations decreases with increased embedment of the pile, due
7.15, 12.05, and 18.60 m, respectively. to increased capacity and a subsequent reduction in reflected
It should be noted that all open-ended piles failed in a fully stress wave from the pile toe. There is also a distinct difference
plugged mode with no change in the measured height of the in the magnitude of the tensile reflection between open- and
soil plug. After the completion of a static test, any subsequent closed-ended piles [compare Figs. 9(a) and 10(a)]. The down-
redriving continued the progression of the soil plug and the ward traveling dynamic force tends to be dissipated by the

FIG. 9. Typical Stress-Wave Data from Open-Ended Pile, Depth 7.15 m (Test J ) Hammer Delivered at Drop Height of 1.8 m: (a) Force
and Zv Data; (b) Integrated Energy and Displacement

FIG. 10. Typical Stress-Wave Data from Closed-Ended Pile, Depth 7.15 m (Test J ) Hammer Delivered at Drop Height of 1.7 m: (a) Force
and Zv data; (b) Integrated Energy and Displacement

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING / NOVEMBER 1999 / 993

J. Geotech. Geoenviron. Eng., 1999, 125(11): 988-998


time the stress wave reaches the base of the pile, due to the diesel hammers. The energy losses arise partly from the impact
additional load transfer from internal shaft friction mobilized (the steel ram striking the pile cap directly) and partly from
along the soil plug height. If the residual stresses locked into the frictional resistance encountered by the ram as it is forced
the pile by driving are large enough, the base resistance will laterally against its bearing by the Coriolis force (de Nicola
be mobilized very rapidly, and no tensile wave will be re- 1996).
flected. However, the residual stresses measured for both piles
were similar and represented only a small proportion of the Effect of Delivered Hammer Energy
bearing capacity.
The pile-head displacements could not be measured directly Different amounts of delivered hammer energy were im-
during the hammer impact, and hence, were calculated by in- parted to the piles by varying the drop height of the ram mass
tegrating the velocity data for the duration of the dynamic from around 0.6 to 1.9 m. At any one pile penetration, up to
perturbation [Figs. 9(b) and 10(b)]. However, the displace- five different drop heights were used in the dynamic tests, with
ments could not be computed beyond a time of about 30 ms, the resulting stress waves measured and used as input into the
owing to reflections emanating from the free end of the Hop- different dynamic analyses. Fig. 11 shows the variation of de-
kinson bar. As such, it was only possible to compute pile-head livered hammer energy with drop height. There is considerable
Downloaded from ascelibrary.org by Aalborg University on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

displacements up to the maximum displacements, which com- scatter, due to varying hammer efficiencies, but a trend of in-
prised both temporary compression and permanent displace- creasing efficiency of the driving system as the hammer drop
ment. The permanent pile sets were taken from the driving height increases. This trend was also observed in a study of
records immediately preceding the static pile test sites, en- dynamic tests conducted on field scale piles by Hussein et al.
abling the magnitude of temporary compression to be esti- (1992), and is perhaps due to certain energy losses that are
mated. This procedure gave values of temporary compression largely independent of drop height.
that were similar to those deduced from (4). An increase in input hammer energy results in an increase
in pile displacements (s and c), and hence, leads to higher
DYNAMIC STRESS-WAVE ANALYSES mobilized soil resistance. This is reflected in Fig. 12, which
shows a clear trend of increasing Rs with increasing drop
Energy Balance Analysis
height.
The simplest approach to estimating the static capacity of The dynamic tests were grouped into four main categories,
the pile (or more correctly, the mobilized soil resistance) is depending on the drop height at which the dynamic tests were
from an energy balance between the input energy from the performed. The static pile capacity estimates from each cate-
hammer and the work done by the pile movement. There is a gory were then compared with the pile capacity measured dur-
variety of pile driving formulas in common use, many of ing the static tests (at a pile-head displacement of 0.1D), with
which have been reviewed by Whitaker (1975). Rather than the results summarized in Table 5. On average, all estimated
describe all of the different approaches here, the basic for- mobilized soil resistances underestimate the measured static
mulation will be outlined and some of the main variables dis- capacity, with the best estimates obtained from dynamic tests
cussed. conducted at 1.5-m drop heights or greater. At low hammer
The basis of the fundamental pile driving formula is rep- drop heights (and hence, low impact energy), the driving for-
resented by mula approach tends to be conservative by around 35% and
improves to within 15% of the static pile capacity at higher
␩/Wh impact energies. Fig. 13 shows how the individual dynamic
Rs = (3)
(s ⫹ c/2) tests, conducted at drop heights of 1.7 m or greater, compare
with the measured static pile capacities. Generally, consider-
where Rs = mobilized soil resistance; ␩ = hammer efficiency; able scatter is present in the calculated soil resistance Rs, with
W = weight of the hammer; h = drop height; s = permanent most of the estimates underpredicting the measured static ca-
set; and c = elastic movement of the pile. The average per- pacities Qs, particularly for values in excess of 12 MN.
manent set was taken from the driving records, whereas the
temporary set was estimated from (de Nicola 1996)
Fp L
c= (4)
(EA)p
where Fp = measured peak impact force; L = length of the
pile; and (EA)p = cross-sectional rigidity of the pile.
The efficiency of a hammer blow ␩ can be determined from
the following equation:

冕 Fv dt
␩= (5)
mgh
where the energy measured during the impact is given by in-
tegrating the product of the pile-head force F and velocity v
over time. The efficiency of the hammer is therefore the ratio
of the energy measured during the impact, to the maximum
possible energy delivered by the hammer of mass m at a drop
height h.
The miniature pile driving hammer performed relatively
well with most hammer blows delivered with average effi-
ciencies of around 60%, which is in a similar range to field FIG. 11. Delivered Hammer Energy as Function of Hammer
scale steam or hydraulic hammers and it is slightly higher than Drop Height

994 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING / NOVEMBER 1999

J. Geotech. Geoenviron. Eng., 1999, 125(11): 988-998


where at a position z along the bar and a time t, the pile
displacement is w and the wave speed is c (Timoshenko and
Goodier 1970).
By solving (6) and deriving the upward and downward com-
ponents of force F from the net force and particle velocity v
at the instrumentation level, an expression for the static pile
capacity Rs is given by (Rausche et al. 1985)

Rs = 0.5(1 ⫺ jc)(F ⫹ Zv)t ⫹ 0.5(1 ⫹ jc)(F ⫺ Zv)t⫹(2L/c) (7)

where jc = Case damping coefficient; and Z = impedance of


the pile. The return time is simply given by 2L/c, which is the
time taken for the stress wave to propagate down the pile
(length L) and return to the instrumentation level.
The deduced mobilized soil resistance Rs can be very sen-
Downloaded from ascelibrary.org by Aalborg University on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

sitive to the value adopted for the damping parameter jc. How-
ever, the above expression can provide useful guidance on the
static pile capacity, where it is possible to calibrate the param-
eter jc for a particular site, either by performing static tests or
by full numerical dynamic analyses of the stress wave data as
described later. Where no static load tests are carried out,
guidelines for jc, as given in Table 6, may be adopted (Rausche
FIG. 12. Estimated Static Pile Capacity Rs for Driving Formula
et al. 1985).
Approach

TABLE 5. Average Ratios ␮ and Standard Deviations ␴ of Cal- Modified Case Analysis
culated to Measured Static Pile Capacity
Eq. (7) must be modified to take into account the large
Hammer drop heights Driving change in impedance that is present in the model piles tested
used in dynamic Formula Case* IMPACT
in the centrifuge (Fig. 14). Consider a downward traveling
pile tests
wave of velocity vd arriving at a point in the pile where the
(m) ␮ ␴ ␮ ␴ ␮ ␴
impedance changes from Z1 to Z2. The incidence wave will
(1) (2) (3) (4) (5) (6) (7)
give rise to a reflected wave with a velocity vu in Region 1,
<1.0 0.66 0.18 0.84 0.19 — — and a transmitted wave with a velocity vd in Region 2. Assum-
1.2 0.71 0.18 0.87 0.23 — —
1.5 0.86 0.20 0.96 0.18 — —
ing that there is no incidence upward traveling wave from
>1.7 0.84 0.18 0.97 0.18 0.98 0.05 Region 2 (which may be treated in an analogous way), the
particle velocity and force at the boundary of the two regions
just after arrival of the downward wave are given by (Ran-
dolph 1990):

v = vincidence ⫹ vreflected = (vd ⫹ vu)1 = (vd ⫹ vu)2 (8)

TABLE 6. Suggested Values for Case Damping Coefficient,


Rausche et al. 1985

Soil type in bearing Suggested range of Correlation value of


strata jc jc
(1) (2) (3)
Sand 0.05–0.20 0.05
Silty sand/sandy silt 0.15–0.30 0.15
Silt 0.20–0.45 0.30
Silty clay/clayey silt 0.40–0.70 0.55
Clay 0.60–1.10 1.10

FIG. 13. Driving Formula Approach, All Test Data for Drop
Heights of >1.7 m

Case Analysis
The governing equation that describes the motion of the
stress wave through an elastic rod in the axial direction is
given by
⭸2w 2 ⭸ w
2

2 = c (6)
⭸t ⭸z2 FIG. 14. Schematic of Change in Pile Impedance

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING / NOVEMBER 1999 / 995

J. Geotech. Geoenviron. Eng., 1999, 125(11): 988-998


F = Z1(vd ⫹ vu)1 = Z1vincidence ⫺ Z1vreflected (9a)
F = Z2(vd ⫹ vu)2 (9b)
From these sets of equations, it may be shown that

v= 冉 2Z1
Z1 ⫹ Z2 冊 vincidence (10)

The corresponding transmitted force, at the impedance change,


is given by

Fd2 = 冉 2Z2
Z1 ⫹ Z2 冊 Fd1 (11)

Fu1 = 冉 2Z1
Z1 ⫹ Z2
冊 Fu2 (12)
Downloaded from ascelibrary.org by Aalborg University on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Eq. (11) and (12) can now be substituted into the standard
Case (7) to yield a modified formulation of the estimated static
pile capacity Rs, which will be referred to as the Case* method,
given by

Rs = 0.5(1 ⫺ jc) 冉 2Z2


Z1 ⫹ Z2
冊 (F ⫹ Zv)t FIG. 16.
>1.7 M
Case* Approach, All Test Data for Drop Heights of

⫹ 0.5(1 ⫹ jc) 冉Z1 ⫹ Z2


2Z1
冊 (F ⫺ Zv)t⫹2L/c
(13)
static pile capacities for dynamic tests conducted at drop
heights of 1.7 m or greater, adopting a value of 0.25 for
jc. This approach yields an average ratio of mobilized soil
Case* Damping Parameter jc resistance to measured static capacities (Rs/Qs) of 0.97, with
As described earlier, the Case* analysis is highly sensitive most estimates lying within ⫾25% of the measured capaci-
to the choice of parameter jc. Because there was no previous ties. As was the case with the driving formula approach, the
experience in stress-wave analysis in silica flour, a study was Case* method underestimates static pile capacities in excess
undertaken to examine the effects of adopting different values of 12 MN.
of jc to the estimated static pile capacity. Once again, the dy-
Numerical Stress-Wave Analysis
namic pile tests were grouped in accordance with the drop
heights used, and a range of jc values was explored in each Predictions of pile capacity for the stress-wave measure-
group, comparing the deduced pile capacities with those mea- ments, using an expression such as the Case* method, is rarely
sured from the static tests. Fig. 15 summarizes the results of used in isolation without calibration either through static load
varying jc that provides the closest approximation to the static tests or by means of a full dynamic analysis of the pile and
capacity varies with the drop height (or impact energy) that is matching of the stress-wave data.
used in the dynamic tests. The process of matching the measured stress-wave data is
In a similar fashion to the driving formula approach, where an iterative one, where the soil parameters for each element
the closest approximations to the static data came from the down the pile are varied until an acceptable fit is obtained
dynamic tests conducted at higher impact energies, a value of between measured and computed results.
0.25 for jc is judged as appropriate for silica flour. Fig. 16 In the numerical analysis, limitations in the simple spring-
shows the correlation between the measured and computed dashpot models for the pile-soil interaction entail that the com-
puter simulation will not match the measured data exactly. A
consequence is that the final distribution of soil parameters
should not be considered as unique, but rather as a best fit
obtained by one operator. Generally, the total static resistance
computed will show little variation among different operators,
or sets of parameters, provided a reasonable fit is obtained,
but the distribution of resistance between shaft and base may
vary (Fellenius 1988).
IMPACT (Randolph 1994) was used to match the stress-
wave data collected in the centrifuge tests. Only the dynamic
data collected from the tests using 1.7-m drop heights or
greater were analyzed with typically good fits obtained be-
tween the measured and calculated stress waves. Where static
pile test data was available, the local shaft friction and end-
bearing pressures were used as the initial input to the numer-
ical analysis. These initial values were often sufficient to pro-
vide a good fit, with only minor adjustments required to
improve the quality of the fit.
Fig. 17 is an example of how the numerical output compares
with measured values for the closed-ended pile previously re-
ferred to in Figs. 10(a and b). Fig. 18 shows the comparisons
between the estimated mobilized soil resistance and measured
FIG. 15. Selection of Case* Damping Parameter jc static pile capacities for each test. Generally, the IMPACT
996 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING / NOVEMBER 1999

J. Geotech. Geoenviron. Eng., 1999, 125(11): 988-998


Regardless of the different failure mechanisms that occur
during static and dynamic loading, the numerical stress-wave
fits from IMPACT managed to estimate the static pile capacity
sufficiently accurately. As was demonstrated by the soil plug
formation during installation (Fig. 7), under dynamic loading
the plug was failing with an IFR close to 1. Instrumented field
scale pile tests conducted by Brucy et al. (1991) have shown
that, under dynamic loading while the pile is in an unplugged
or partially plugged mode, the contribution of internal shaft
resistance mobilized along the soil plug is only a small pro-
portion of the total mobilized soil resistance. Instrumented pile
tests conducted at model scale by Paik ad Lee (1993) and
O’Neill and Raines (1991) also confirm this observation.
On average, the piles were permanently displaced around
0.04D at the shallow test sites, reducing to 0.02D at the ulti-
Downloaded from ascelibrary.org by Aalborg University on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

mate penetration depths. It is surprising that while the piles


were displaced <0.1D during a dynamic blow (which is the
reference displacement for the static bearing capacity), the mo-
bilized soil resistance was comparable with the static pile ca-
FIG. 17. Calculated (IMPACT) versus Measured Stress-Wave pacity, even though the major proportion of the static capacity
Data emanated from the pile base.
A partial explanation may be that the amount of shaft re-
sistance mobilized along the plug height during loading com-
pensates for the incomplete mobilization of end bearing pres-
sure. As a result, the estimation of total soil mobilization from
the dynamic data corresponds well to that measured at greater
pile-head displacements during static loading.

CONCLUSIONS
Dynamic stress-wave data measured from centrifuge pile
tests have been used as input in three common methods of
dynamic analysis. Estimations of the static pile capacity (or
mobilized soil resistance) were derived from the dynamic anal-
yses and compared with the measured pile capacities from
static tests. It was evident that the dynamic tests performed
with the highest input energy provided the most accurate es-
timations of the static pile capacity because greater pile-head
displacements (and hence soil mobilization) were achieved.
The simplest approach was based on a fundamental energy
balance and it provided reasonably good approximations, typ-
ically underpredicting the static pile capacity by 15%. Pile
capacity estimates from the energy balance method were
highly dependent on input energy and provided less accurate
FIG. 18. IMPACT Analysis, All Test Data estimations as the measured static pile capacity increased.
A modified Case method, which incorporated large changes
analysis provides an excellent estimation to the static capacity in pile impedance, was also applied to the dynamic data. A
with the ratio of Rs /Qs = 0.98. With a standard deviation of value of 0.25 for the damping parameter jc provided the best
0.05, there is considerably less scatter in the estimates com- approximation to the measured static capacity and also com-
pared to both the Driving Formula and Case* analyses, which pared well with the suggested value from Rausche et al.
both had standard deviations of 0.18. As discussed earlier, a jc (1985). This value of jc was determined from the dynamic
value in the range of 0.15–0.25 provided the best fit, with the tests, conducted at the highest delivered energy, and calibrated
capacities deduced either from the IMPACT analyses or the with the static pile test data.
static load tests. Of the three dynamic analyses, the numerical analysis using
IMPACT provided, consistently, the best estimates for static
FAILURE MODES DURING STATIC AND pile capacity. This was expected because considerable effort is
DYNAMIC LOADING needed to determine a suitable profile of soil resistance along
It is interesting to consider the different failure mechanisms the pile shaft and base to match the measured and computed
that occur in pipe piles during dynamic and static loading. As stress waves. The deduced total soil resistance therefore tends
mentioned earlier, it is generally accepted that a pile remains to give an accurate prediction of the static pile capacity.
substantially fully plugged in the static loading state, but be- The resulting good agreement between pile capacity de-
comes unplugged under the dynamic impact loading condition. duced from dynamic and static load testing must still be
For a fully plugged pipe pile, sufficient inside shaft resistance viewed in the light of significant differences between the two
can be mobilized between the pile and the soil plug, and hence, types of test, particularly for open-ended piles. In the dynamic
an equivalent end-bearing capacity over the full cross-sectional test, the pile movement is much less than the static displace-
area is considered. For an unplugged pile, the integrated inside ment at which the capacity is quoted. Additionally, failure for
shaft resistance must be less than the sum of the end-bearing open-ended piles occurs in a (largely) unplugged mode, com-
resistance of the plug and the inertial resistance of the plug pared with a fully plugged mode during static testing. The
(Stevens 1988). good agreement therefore implies that, during continuous driv-
JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING / NOVEMBER 1999 / 997

J. Geotech. Geoenviron. Eng., 1999, 125(11): 988-998


ing the total internal shaft friction maintains a close balance ing at a function of ram drop height.’’ Proc., 4th Int. Conf. on Appl.
with the (potential) gross end-bearing resistance, giving com- Stress Wave Theory to Piles, 421–424.
Jardine, R. J., Lehane, B. M., and Everton, S. J. (1992). ‘‘Friction coef-
parable dynamic unplugged, or static plugged, capacities. ficients for piles in sands and silts.’’ Proc., Int. Conf. on Offshore Site
Investigation and Found. Behaviour, Soc. of Underwater Technol.,
ACKNOWLEDGMENTS 661–680.
Kishida, H., and Uesugi, M. (1987). ‘‘Tests of the interface between sand
The work described in this paper forms part of an extensive research and steel in the simple shear apparatus.’’ Géotechnique, London, 37(1),
program into driven piles in sand, in the Special Research Centre for 45–52.
Offshore Foundation Systems at the University of Western Australia, es- Maslen, C. (1997). ‘‘Constant normal stiffness testing of silica soils.’’
tablished and supported under the Australian Research Council’s Re- Undergraduate Honours Project, University of Western Australia, Perth,
search Centres Program. D. Bruno holds a university postgraduate award Australia.
and a William Lambden Owen supplementary scholarship funded by the O’Neill, M. W., and Raines, R. D. (1991). ‘‘Load transfer for pipe piles
University of Western Australia. This financial support is gratefully ac- in highly pressured dense sand.’’ J. Geotech. Engrg., ASCE, 117(8),
knowledged. 1208–1226.
Paik, K. H., and Lee, S. R. (1993). ‘‘Behaviour of soil plugs in open-
ended model piles driven into sands.’’ Marine Georesources and Geo-
APPENDIX. REFERENCES
Downloaded from ascelibrary.org by Aalborg University on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

technology, 11, 353–373.


Randolph, M. F. (1990). ‘‘Analysis of the dynamics of pile driving.’’
Barends, F. B. J. (1992). Proc., 4th Int. Conf. on Application of Stress-
Developments in soil mechanics—IV: Advanced geotechnical analysis,
Wave Theory to Piles.
P. K. Banerjee and R. Butterfield, eds., Elsevier Science.
Broms, B. B., ed. (1981). Proc., 1st Int. Conf. on Application of Stress-
Randolph, M. F. (1991). ‘‘Establishing a new centrifuge facility.’’ Proc.,
Wave Theory to Piles.
Int. Conf. Centrifuge 91, 1, 3–9.
Broms, B. B., ed. (1985). Proc., 2nd Int. Conf. on Application of Stress-
Randolph, M. F. (1994). ‘‘IMPACT: Dynamic analysis of pile driving.’’
Wave Theory to Piles. User Manual, University of Western Australia, Perth, Australia.
Bruno, D. (1998). ‘‘Centrifuge modeling of driven piles in dense sand.’’ Rausche, F., Goble, G. G., and Likins, G. E. (1985). ‘‘Dynamic deter-
Proc., 3rd ANZ Young Geotech. Profl. Conf., 9–14. mination of pile capacity.’’ J. Geotech. Engrg., ASCE, 111(3), 367–
Bruno, D., and Randolph, M. F. (1998). ‘‘Dynamic testing of driven piles 383.
in dense sand.’’ Proc., Int. Conf. CENTRIFUGE 98 (ISSMFE Centri- Schofield, A. N. (1980). ‘‘Cambridge geotechnical centrifuge opera-
fuge 98). tions.’’ Géotechnique, London, 30(3), 227–268.
Brucy, F., Meunier, J., and Nauroy, J. F. (1991). ‘‘Behaviour of pile plug Smith, E. A. L. (1960). ‘‘Pile driving analysis by the wave equation.’’ J.
in sandy soils during and after driving.’’ Proc., 23rd Annu. Offshore Soil Mech. and Found. Div., ASCE, 86, 35–61.
Technol. Conf., 145–154. Stevens, R. F. (1988). ‘‘The effect of a soil plug on pile drivability in
de Nicola, A. (1996). ‘‘The performance of pipe piles in sand,’’ PhD clay.’’ Proc., 3rd Int. Conf. on Applied Stress Wave Theory to Piles,
thesis, University of Western Australia, Perth, Australia. 861–868.
Fellenius, B. H., ed. (1988). ‘‘Variation of CAPWAP results as a function Timoshenko, S. P., and Goodier, J. N. (1970). Theory of elasticity. 3rd
of the operator.’’ Proc., 3rd Int. Conf. on Application of Stress-Wave Ed., McGraw-Hill, New York.
Theory to Piles, 814–825. Townsend, F. C., ed. (1996). Proc., 5th Int. Conf. on Appl. Stress Wave
Fleming, W. G. K., Weltman, A. J., Randolph, M. F., and Elson, W. K. Theory to Piles.
(1992). Piling engineering 2nd Ed., Black Academic & Professional. Whitaker, T. (1975). The design of piled foundations. 2nd Ed., Pergamon,
Goble, G., and Likins, G. (1996). ‘‘On the application of PDA dynamic Oxford. Tarrytown, N.J.
pile testing.’’ Proc., 5th Int. Conf. on the Application of Stress-Wave Zelikson, A. (1991). ‘‘Transient signals from below-ground transducers
Theory to Piles, 263–273. during pile-driving and static loading: a hydraulic gradient simulation
Hussein, M., Rausche, F., and Likins, G. (1992). ‘‘Dynamics of pile driv- study.’’ Géotechnique, London, 41(4), 553–569.

998 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING / NOVEMBER 1999

J. Geotech. Geoenviron. Eng., 1999, 125(11): 988-998

You might also like