Sei sulla pagina 1di 14

1142

Axial testing and numerical modeling of square


shaft helical piles under compressive and tensile
loading
Ben Livneh and M. Hesham El Naggar

Abstract: Helical piles are increasingly used to support and rehabilitate structures subjected to both tensile and compressive axial loads. This paper presents a detailed investigation into the axial performance of helical piles. The study encompasses 19 full-scale load tests in different soils and numerical modeling using finite element analysis. The ultimate load
criteria and load transfer mechanisms for helical piles were examined. In addition, the relationship between the installation
effort (torque) and pile capacity was explored to determine its suitability for predicting pile capacity. The piles tested were
made of three circular pitched bearing plates welded at a spacing of three helical diameters to a solid-square, slender steel
shaft. It is proposed to determine the ultimate pile capacity as the load corresponding to pile head movement equal to 8%
of the largest helix diameter plus the pile elastic deflection. A torque correlation factor, KT = 33 m1 for compression and
KT = 24 m1 for uplift, was established to relate the ultimate pile capacity to the installation torque. It was found that load
transfer to the soil is predominantly through a cylindrical shear failure surface that follows the tapered profile of the interhelices soils and the bearing capacity of the lead helix in the direction of loading.
Key words: helical screw piles, load transfer mechanism, failure criterion, individual bearing, cylindrical shear, torque correlation.
Resume : Des pieux helicodaux sont utilises de plus en plus pour soutenir et rehabiliter des structures soumises a` des
chargements axiaux tant en traction quen compression. Cet article presente une etude detaillee sur la performance axiale
des pieux helicodaux. Letude comprend 19 essais de chargement a` pleine echelle dans differents sols, et une modelisation
numerique au moyen dune analyse en elements finis. On a examine le crite`re de chargement ultime et les mecanismes de
transfert de chargement des pieux helicodaux. De plus, la relation entre leffort dinstallation (torque) et la capacite du
pieu a ete exploree pour determiner sa pertinence pour predire la capacite du pieu. Les pieux testes etaient constitues de
trois plaques circulaires portantes en forme de vis soudees a` un espace de trois diame`tres dhelice a` un essieu carre-solide
elance en acier. On propose de determiner la capacite ultime du pieu comme etant la charge correspondant au mouvement
de la tete egal a` 8 % du diame`tre de lhelice le plus grand plus la deflexion elastique du pieu. Un facteur de correlation
du torque, KT = 33 m1 pour la compression et KT = 24 m1 pour larrachement a ete etabli pour mettre en relation la capacite ultime du pieu avec le torque dinstallation. On a trouve que le transfert de charge au sol se fait de facon predominante le long dune surface de rupture en cisaillement cylindrique qui suit le profil conique des sols entre les helices; et la
capacite portante de lhelice de tete dans la direction du chargement.
Mots-cles : pieux en forme de vis helicodale, mecanisme de transfert de charge, crite`re de rupture, portance individuelle,
cisaillement cylindrique, correlation de torque.
[Traduit par la Redaction]

Introduction
Helical piles (also referred to as anchors, anchor piles, or
screw piles) have most commonly been used as anchors,
to resist tensile loads in supporting structures such as lighthouse beacons, buried pipelines, utility poles, guyed towers,
and transmission towers. In recent decades, their applications in engineering projects have expanded to both support
and rehabilitate structures under tensile, compressive, and
Received 3 November 2007. Accepted 30 March 2008.
Published on the NRC Research Press Web site at cgj.nrc.ca on
31 July 2008.
B. Livneh and M.H. El Naggar.1 Geotechnical Research
Centre, Faculty of Engineering, The University of Western
Ontario, London, ON N6A 5B9, Canada.
1Corresponding

author (e-mail: helnagg@engga.uwo.ca).

Can. Geotech. J. 45: 11421155 (2008)

lateral loading. Given their current increase in application,


however, the amount of available research and design methodologies to date are relatively limited in comparison with
other, more conventional piling solutions, and the majority
of relevant research has been focused solely on the uplift
condition. Therefore, given the lack of design procedures,
the application of helical piles, particularly for the compression case, is considered in this investigation.
The combination of variable shaft length and variable helix diameter has expanded the range of projects for which
helical piles may be suitable. Their installation method allows them to reach a great depth with the addition of extension segments in the field, and thus increases their
versatility. Helical piles are installed almost vibration-free,
through the use of mechanical torque, which reduces damage to adjacent structures, and they can be constructed without excavating soil or pouring concrete. This allows for

doi:10.1139/T08-044

2008 NRC Canada

Livneh and El Naggar

cleaner and quicker installation and makes them both environmentally friendly and cost effective.
Early approaches towards the evaluation of helical anchor
capacity involved examining the behaviour of shallow single
plate anchors. Utilizing either assumed or observed failure
surfaces within the soil adjacent to the pile, several studies
characterize the failure geometry for anchors. Majer (1955),
Mors (1959), and Ireland (1963) proposed highly idealized
conical (reaching the surface at an angle of 908 + , where
 is internal friction angle of the soil) and cylindrical failure
surfaces. More recently, theoretical studies on anchors conducted by Vesic (1971), Rowe and Davis (1982), and
Saeedy (1987) included limit equilibrium and finite element
analyses, which most notably focused on the behaviour of
the soil immediately contiguous to the anchor and on
whether its state of stress provided evidence of resistance to
anchor loading.
Laboratory investigations of anchor behaviour and failure
geometry include half-scale and full-scale laboratory models
by Balla (1961), Sutherland (1965), Downs and Chieurzzi
(1966), Meyerhof and Adams (1968), Clemence and Veesaert (1977), Sutherland et al. (1982), Murray and Geddes
(1987), Weizhi and Fragaszy (1988), Hoyt and Clemence
(1989), Ghaly and Hanna (1991), Ghaly et al. (1991), and
Narasimha Rao et al. (1993). The results included estimations of a failure surface reaching the ground surface at angles between /4 and /2 to the vertical. The behaviours of
shallow and deep anchors were classified through either a
failure surface extending to the ground surface (shallow anchor behaviour), or a localized shearing failure (deep anchor
behaviour). Based on laboratory results, Balla (1961) established a breakout factor as a dimensionless quantity related
to the peak pullout load (H/D: where H is the depth of
embedment of the uppermost helix and D is the diameter
of the largest helix), which can also be used to classify shallow and deep anchors.
Narasimha Rao et al. (1989) conducted an experimental
program with multihelix anchors showing that pile ultimate
uplift capacity increases with (i) the number of helical
plates; (ii) decreasing soil moisture content; and (iii) increasing soil consistency index. The development of a cylindrical
failure surface below the top helix was shown for piles with
small helical spacing (i.e., S/D 3, where S is the space between helical plates and D is the average helical diameter).
In this regard, the results of Narasimha Rao and Prasad
(1991) and Narasimha Rao et al. (1993) were consistent
with the findings of Mitsch and Clemence (1985), who presented results of both laboratory and field investigations on
triple helix anchors. Mitsch and Clemence (1985) provided a
method for estimating the uplift capacity of shallow and
deep piles, dependent on pile embedment, helical spacing,
and soil conditions.
Aside from the empirical estimation of helical pile capacity through a correlation to installation torque, there presently exists two general theories describing the failure
mechanism of multihelix anchors, namely through cylindrical shearing, involving the development of a failure surface
between the interhelical soil, and through individual bearing
of each helical plate, where each helix behaves independently. The distinction between these methods has significant

1143

implications on pile ultimate capacity and is of particular interest for this investigation.

Experimental investigation
The main objective of this research is to evaluate the axial
performance of helical piles by means of a full-scale loading
program and the development of a numerical model. The
specific objectives of the full-scale load testing program
were (i) to determine the load transfer mechanism for compressive and tensile loading in different ground conditions;
(ii) to define appropriate ultimate load criteria for helical
piles; (iii) to evaluate the compressive and tensile capacity
of tested helical piles based on these criteria; and (iv) to explore the relationship between the installation effort (torque)
and the capacity of the pile and determine whether this relationship can be used as a predictor for the pile capacity and
(or) performance acceptance of the pile. The numerical
model was used to further address some of the research objectives, namely, to determine the load transfer mechanism
and to establish a theoretical model to evaluate the capacity
of helical piles under different loading conditions.

Pile description
The SS175 Chance foundation system used in this evaluation is manufactured by AB Chance Company (Centralia,
Mo) and it consists of three helical bearing plates (diameters
300 mm, 250 mm, and 200 mm decreasing with depth)
welded to a central shaft (44.5 mm) and it is defined as a
segmented deep foundation system. Extension segments (or
sections) are attached to the lead section with bolted couplings during installation to allow the system to bear upon
soil at a desired depth. The helical shape of the bearing
plates allows for minimum soil disturbance during installation and because each helix is a single 75 mm pitch of a
screw thread, the system can literally screw into the ground.
The path of each consecutive helix follows the same path as
the preceding one during installation in such a scenario.
The lead section of the pile (Fig. 1) supports the loads applied to the system by transferring them to the soil. The
spacing between the plates is approximately three times the
diameter of the lower plate (i.e., S/D & 3, or 750 mm and
600 mm with increasing depth), and it provides improved
pile capacity as this preferred spacing is prescribed as the
interval between the two governing failure mechanisms,
namely individual bearing and cylindrical shearing (Hubbell
Power Systems Inc. 2003).

Site investigation, pile installation, and


loading apparatus
The load testing program was conducted at two locations
on the Environmental site at The University of Western Ontario, London. Two boreholes were performed using a hollow stem auger mounted on a rig. The soil profiles are
composed of layers of clayey and sandy silts overlying deep
deposits of either stiff silty clay (site 1) or dense-fine sand
(site 2). The ground water table elevation in the site varied
between 5.2 and 6.5 m below the ground surface (based on
these boreholes and other boreholes in the site area performed before and after this testing program). The standard
#

2008 NRC Canada

1144

Can. Geotech. J. Vol. 45, 2008

Fig. 1. Schematic of a typical pile lead section. f, internal friction angle of the soil.

Table 3. Summary of geotechnical properties of soil from


site 1 at 2.4 m.

Table 1. Borehole and SPT results from site 1.


Layer
Stiff brown sandy clayey silt
Very stiff brown clayey silt (W.T. depth)
Stiff grey clayey silt
Very stiff grey sandy clayey silt
Dense grey silt

Depth (m)
2.4
4.1
5.8
7.3
>7.3

N-value
19
25
11
17
30

Note: W.T., water table; N-value, SPT value (i.e., number of blows).

Table 2. Borehole and SPT results from site 2.


Layer
Stiff brown sandy clayey silt
Very stiff grey clayey silt (W.T. depth)
Very stiff grey clayey silt, embedded
gravel
Dense grey fine sand, trace of silt

Depth (m)
2.6
4.1
5.2

N-value
34
21
20

>5.2

32

Note: W.T., water table; N-value, SPT value (i.e., number of blows).

penetration test (SPT) was conducted at each site to a minimum depth of 8 m, using a safety hammer with a rope and
cathead. Advancing a split spoon sampler, the SPT results
are provided in Table 1 and Table 2. Three consolidated undrained triaxial tests were conducted on spilt spoon soil
samples retrieved at a depth of 2.4 m at site 1, to evaluate
the total and effective shear strength parameters (undrained
shear strength, cu, cohesion, c, and angle of internal friction,
). The results of this test along with the combined results
of sieve and hydrometer analysis from this location are provided in Table 3.
Complete logs of pile installation included readings of
torque, taken at depth intervals of 0.3 m (1 ft). The profiles
of installation torque throughout the depth of installation are
shown in Fig. 2. Based on an empirical correlation factor,
KT, the installation torque readings were used to estimate
pile capacity, such that Qt = KTT (where Qt is the predicted
pile capacity and T is the installation torque of the pile). The
estimated capacity was used to prescribe the load increments
in the testing program. This correlation has long been used
in the field, with the rationale that installation torque is a
measure of the energy required to overcome the shear
strength of the soil and is hence directly related to pile capacity.
During installation of piles to be tested in compression,

Gravel content (%)


Sand content (%)
Silt-clay content (%)
Specific gravity
Moisture content (%)
Liquid limit
Plastic limit
Plasticity index
Undrained shear strength, Su = cu (kPa)
Undrained elastic modulus, Eu (MPa)
Effective strength, c (kPa)
Internal friction angle,  (8)
Drained elastic modulus, E (MPa)

1.40
62.80
35.8 (38% clay
and 62% silt)
2.85
23.50
29
25
4
60
60
10
28
54

the annular cavity (~90 mm) created by the square pile-shaft


and the broad cross-section of the forged socket of the coupler means was constantly filled with a cementitious-based
grout. After curing, the grout effectively increased the buckling resistance of the slender shaft by mobilizing the surrounding soil resistance, as well as provided additional
rigidity to the connections. Reaction piles were installed at
a distance of 2.7 m from the testing piles at roughly the
same depth as the testing pile to effectively anchor the testing frame.
Figure 3a illustrates the general layout for the setup for
compression load testing. The figure shows the reaction
frame, which consists of four reaction piles (in tension), the
load beam (steel 600  242 I-beam), and two spreader
beams (two steel 300  88 I-beams). The pile head was
fixed using a 150 mm diameter hollow PVC pipe that extended approximately 1.3 m below the ground, around
which was secured a 1 m T-pipe that was grouted and
served as the pile cap; a flat, leveled surface for the loading
jack. The load was measured using a calibrated load cell between the hydraulic jack and the loading frame and also
through the measurements recorded on the hydraulic pressure of the loading jack. The apparatus for tension load testing is shown in Fig. 3b. The load readings were taken in a
similar manner to the compression testing. For all cases, the
displacements were monitored using three dial-gauges configured on the pile cap.
#

2008 NRC Canada

Livneh and El Naggar


Fig. 2. (a) Installation torque versus depth for piles at site 1. (b)
Installation torque versus depth for all piles at site 2. (c) Installation
torque versus depth for piles adjacent to site 1.

1145

The axial pile load tests were conducted according to


ASTM D114381 (ASTM 1994) and ASTM D368990
(ASTM 1995) in which loads were applied in increments of
approximately 10% of the estimated pile capacity in 5 min
time intervals. In some cases, loading increased beyond the
estimated pile capacity until continuous jacking was required to maintain the load or continuous creeping settlement was observed (i.e., failure).

Load testing program


In total, 19 test piles were installed in two separate sites
in the field. Seven of the test piles at site 1 were installed
such that the lead sections rested within the upper clayey
silt layers (approximately 5 m depth), and seven piles were
installed such that the lead sections rested within the lower
dense silt layer (approximately 9 m depth). At site 2, one
pile was founded in clayey silt and four piles were installed
such that all lead sections lay in sandy soil at a maximum
depth of 7.8 m. According to the definitions proposed by
Ilamparuthi et al. (2002) and by Meyerhof and Adams
(1968), all piles tested in this evaluation are classified as
deep piles based on their embedment depths and would
hence experience a localized failure surface that does not
extend to the ground surface.

Load and deflection data


The pile capacity was estimated knowing its installation
torque using a nominal value of KT = 33 m1 (10 ft1) for
compression piles and KT = 26 m1 (8 ft1) for tension piles,
based on the results of previous investigations (e.g., Hoyt
and Clemence 1989). The load was applied in a minimum
of 10 increments. For each load increment, the readings
from the three dial gauges were taken and averaged.
The incremental readings from the load cell were plotted
against the averaged dial gauge readings to yield a load
deflection curve for each pile tested. These curves provided
a useful tool to establish the ultimate pile capacity. The ultimate failure load for a pile may be defined as the load when
the pile plunges, or settlement increases rapidly under sustained load (Prakash and Sharma 1990). Plunging failure
typically involves settlements that far exceed the acceptable
range for design and may not always be attained during a
load test. If plunging does not occur clearly, another definition of ultimate load is needed.
The loadmovement curves for piles generally include
three different regions: an initial linear region with large
slope (i.e., high stiffness); a strongly nonlinear region with
pile movements disproportionately larger for each load increment; and finally, a nearly linear region with small slope
(i.e., low stiffness).
The onset of failure in the loadmovement curve is typically found within the nonlinear portion of the curve,
between the initial linearelastic portion and the final, nearlinear rapid failure section. It is within this region that the
ultimate pile capacity is defined under most failure criteria.
The special case of plunging failure describes a situation
where the final, rapid-failure region of the loadmovement
curve is nearly horizontal (i.e., slope & 0). The ultimate
load of the pile is taken to be the load that occurs at the onset of plunging for this special case.
#

2008 NRC Canada

1146

Can. Geotech. J. Vol. 45, 2008

Fig. 3. (a) Depiction of load testing apparatus for compressive load application; beam positioning in relation to test pile and wood cribbing.
(b) Depiction of load testing apparatus for tensile load application; beam positioning in relation to test pile and wood cribbing.

Failure criterion applied to load test results


Based on the pile loadmovement trends, it is noted that
the behaviour was different for piles tested in tension and
compression. The pile behaviour in the uplift case was initially very flexible, attributed to overcoming the slack of
the loading apparatus (i.e., the reaction beam supported by
wood cribbing, bearing on adjacent, relatively loose surficial
soil). After undergoing large initial displacements of between 2 and 10 mm associated with the first load increment,
these piles exhibited a curvilinear tendency until failure ensued, resulting in large displacements associated with each
successive load increment in the nonlinear region. This trend
was particularly pronounced for piles tested in clayey silt.
Thus, it was necessary to exclude the initial slack displacement from the estimation of a failure criterion since
this behaviour was the result of the testing apparatus and
not the pile itself. For compressive loading, the initial
linearelastic region started from the beginning of the load
test. Collectively, the near-linear failure region was achieved
at a net displacement of greater than 8% of the largest helical diameter (i.e., 0.08D).
Historically, estimates of pile capacity have been made
considering various schemes related to pile geometry and
the shape of the loaddeflection curve, as depicted in Table 4. Most investigators consider some fraction of the characterstic pile dimension D (width for shallow foundations;

Table 4. Commonly used failure criteria for interpreting pile capacity (adapted from Zhang et al. 2005).
Failure criterion
AS-2159 (SAA 1995)

Davissons criterion
(Davisson 1972)
FDOT criterion (FDOT
1999)
FHWA criterion (Reese and
ONeill 1988)
ISSMFE criterion (ISSMFE
1985) and BS 8004
criterion (BSI 1986)
Slope and tangent method
(Butler and Hoy 1977)

Displacement at failure
50 mm at 1.5 times the design
load and 30 mm at unloading
15 mm at serviceability load
and 7 mm upon unloading
PL
D
AE 120 4 mm
PL
AE

D
30
for piles with D > 0.61 m

5%D
10%D

Defined as the deflection at the


intersection of tangents to the
linear-elastic section and plunging failure section*

*Slope of plunging section to be equal to 1 in./40 kip (14.3 mm/100 kN).

diameter for piles), elastic displacement, or intersection of


loaddeflection tangents to define a failure criterion. Thus,
given the unique geometry of helical piles, the failure criterion used to describe the amount of settlement experienced
at the pile head at the ultimate pile load was related to the
#

2008 NRC Canada

Livneh and El Naggar

largest
 helical diameter, D, and elastic deflection of the pile
PL
AE , such that:
1

PL
0:08D
AE

1147
Fig. 4. (a) Tensile loaddeflection curve for a pile tested in clayey
silt. (b) Compressive loaddeflection curve for a pile tested in
dense silt. (c) Compressive loaddeflection curve for a pile tested
in sand.

where S is the settlement experienced at failure; P is the applied load at failure; L is the length of the pile; A is the
cross-sectional area of the pile shaft; E is the Youngs modulus for the steel; and D is the diameter of the largest helix.
Figure 4a shows the loaddeflection curve and failure criterion for a pile tested in clayey silt in tension. The shape of
this graph was characteristic of piles tested in tension and
the failure criterion consistently fell within the nonlinear region for these curves. The loaddeflection behaviour and
failure criterion for a pile tested in dense silt in compression
is presented in Fig. 4b. The plunging behaviour exhibited by
this pile was exclusive to compression piles in dense silt.
The typical loaddeflection behaviour for nonplunging compression piles is observed in Fig. 4c. It is noted that the initial-linear region of load deflection curves for piles tested in
clayey silt exhibited very steep slopes relative to other piles,
while piles tested in sand showed a relatively steep, curvilinear tendency throughout the loading cycle. For all piles, the
application of the proposed failure criterion led to the conclusion that piles installed in sand had the greatest ultimate
capacity followed by dense silt and clayey silt. Additionally,
compression piles carried greater loads than tension piles;
and within the same soil layer, piles installed to a greater
depth had higher ultimate capacities than shallower ones.

Installation torque ultimate pile capacity


relationship
Various authors have confirmed a relationship between installation torque and ultimate pile capacity (Hoyt and Clemence 1989; Ghaly and Hanna 1991). Based on the load test
results, Tables 5 and 6 summarize the loadtorque correlation, KT, for compression and uplift, respectively. The bestfit line was translated to intercept the origin, while preserving
the slope, yielding proposed values that are inherently conservative. The results from two load tests were not representative of the typical loadtorque relationship: pile No. 3 due
to the slight eccentricity caused by a misalignment of the
loading system and pile No. 18 due to severe damage of its
bottom helix caused by striking a cobble (which was visually
confirmed upon retrieving the pile).
Table 5 presents a ratio of installation torque to ultimate
compressive capacity, KT, of between 35.3 m1 (10.6 ft1)
and 62.1 m1 (18.9 ft1), excluding pile No. 3. The values of
KT in dense and clayey silt range between 35.3 m1 (10.6 ft1)
and 42.5 m1 (13 ft1), while in sand the value is greater than
60 m1 (18.2 ft1).
The value of KT for tension piles varied between 21.3 m1
(6.5 ft1) and 36.3 m1 (11.1 ft1) as shown in Table 6. The
upper range values correspond to piles tested in dense silt
and sand and the lower range belongs to piles tested in
clayey silt. The representative torque constants for piles
tested in sand lies between 24.3 m1 (7.4 ft1) and 32.7 m1
(10.0 ft1). These values are in agreement with the findings
reported in the literature.
Figure 5 illustrates the relationship between ultimate com-

pressive capacity and the torque averaged over the final 1 m


of installation. The curve confirms that the pile capacity is
directly proportional to the installation torque and the slope
of the curve represents the torque correlation constant, KT.
Figure 6 shows the relationship between pile ultimate uplift
capacity and the average installation torque over the final
1 m of installation. The curve shows a direct relationship
#

2008 NRC Canada

1148

Can. Geotech. J. Vol. 45, 2008


Table 5. Pile capacity and torque constants for piles tested in compression.
Pile
No.
1
2
3
4
5
6
13
14

Site
No.
1
1
1
1
1
1
2
2

Depth, D,
of top helix
(m) [ft]
7.6 [25]
7.6 [25]
7.6 [25]
3.4 [11]
3.0 [10]
3.0 [10]
5.8 [19]
6.1 [20]

Soil type
Dense silt
Dense silt
Dense silt
Clayey silt
Clayey silt
Clayey silt
Sand
Sand

Pile capacity
(kN) [kip]
465 [104.6]
490 [110.2]
360 [80.9]
350 [78.7]
320 [72]
325 [73]
695 [156.2]
660 [148.4]

Average torque
over last 1m
(kNm) [lbft]
11.76 [8666]
11.54 [8500]
13.80 [10166]
9.95 [7333]
8.73 [6433]
8.60 [6333]
11.31 [8333]
10.63 [7833]

KT
(m1) [ft1]
39.5 [12]
42.5 [13]
26.1 [8]
35.3 [10.6]
36.5 [11.1]
37.8 [11.5]
61.5 [18.7]
62.1 [18.9]

Note: Pile capacity observed at 8%D plus elastic displacement.

Table 6. Pile capacity and torque constants for piles tested in tension.
Pile
No.
7
8
9
16
20
10
17
18
15
12
19

Site
No.
1
1
1
1a
2a
1
1b
1b
1a
2
2a

Depth, D,
of top helix
(m) [ft]
3.7 [12]
4.0 [13]
3.0 [10]
3.0 [10]
2.1 [7]
4.3 [14]
8.2 [27]
8.5 [28]
9.1 [30]
5.2 [17]
7.9 [26]

Soil type
Clayey silt
Clayey silt
Clayey silt
Clayey silt
Clayey silt
Clayey silt
Dense silt
Dense silt
Dense Silt
Sand
Sand

Pile capacity
(kN) [kip]
190 [42.8]
260 [58.3]
160 [36.0]
180 [40.5]
295 [66.2]
245 [55.0]
220 [49.5]
165 [37.0]
350 [78.7]
300 [67.4]
360 [80.9]

Average torque
over last 1.5 m
(kNm) [lbft]
8.91 [6570]
9.80 [7200]
6.55 [4800]
7.64 [5600]
10.10 [7400]
9.93 [7300]
7.07 [5200]
7.64 [5600]
9.63 [7100]
12.35 [9100]
10.99 [8100]

KT
(m1) [ft1]
21.3 [6.5]
26.5 [8.1]
24.4 [7.5]
23.6 [7.3]
29.2 [9.0]
24.7 [7.6]
31.1 [9.5]
21.6 [6.6]
36.3 [11.1]
24.3 [7.4]
32.7 [10.0]

Note: Pile capacity observed at 8%D plus elastic displacement.

similar to that for compression piles. Plotted on the same


scale, it yields however, a slightly smaller value of KT. For
design purposes, the solid lines shown in Fig. 5 and Fig. 6
are proposed.
Consistent with previous findings, installation torque
averaged over the last 3 ft of installation was shown to be
correlated with ultimate pile capacity. Beyond the inherent
uncertainty of the condition of the soil beneath the lowest
helix (compression case), this value provides a reasonable
estimate of the strength of the soil surrounding the pile. Altogether, KT was found to be dependent on both soil strength
(Fig. 2) as well as on loading direction. Values of KT tended
to be higher (lower) in compression (tension) and varied directly with soil strength.

Further analysis
Several of the piles tested in this investigation were instrumented with strain monitoring equipment. The results of
the strain monitoring were limited to a great extent by the
deleterious effects of pile installation. However, strain data
obtained during load testing was used to verify load transfer
in the development of a finite element model (FEM).

Finite element model development


The main objective of modeling helical pile behaviour
was to define the failure mechanism and loadtransfer be-

haviour for each pile. Upon calibrationverification with the


experimental data, the FEM provided insight into the effects
of pile loading on the surrounding soil. Based on the findings of the model and full-scale load test results, a methodology for calculating pile capacity was developed.
To account for the unique geometry of the problem a software program capable of realistically analyzing threedimensional soilfoundation interaction, namely the Plaxis
3D Foundation suite, was selected. To minimize boundary
effects on pile responses, the lateral boundaries of the numerical model were placed at 4 m from the pile centre and
the vertical (bottom) boundary was placed at 40 m below
the ground surface. This allowed for a buffer of approximately 13 pile diameters (of largest helix) between the pile
and any lateral boundary, and 30 m, or approximately 3 pile
lengths in the vertical direction. The boundary conditions
were as follows:
(1) Vertical model boundaries with their normal in x-direction
(i.e., parallel to the yz plane) are fixed in the x-direction
(ux = 0) and free in y- and z-directions.
(2) Vertical model boundaries with their normal in z-direction
(i.e., parallel to the xy plane) are fixed in the z-direction
(uz = 0) and free in x- and y-direction.
(3) Vertical model boundaries with their normal neither in
x- nor in z-direction (skewed boundary lines in a work
plane) are fixed in the x- and z-directions (ux = uz = 0)
and free in the y-direction.
#

2008 NRC Canada

Livneh and El Naggar


Fig. 5. Pile compressive capacity versus installation torque.

1149
Fig. 7. Distribution of wedge elements within the model and nodes
and stress points in a typical 15-node wedge element. z, x, horizontal plane; h, vertical dimension.

Fig. 6. Pile uplift capacity versus installation torque.

grouted column associated with compression piles was


ignored in the model, given the limited resistance provided,
as well as its variable consistency in the field. The models
were validated through a comparison with measured pile responses and strains along the pile during the load test.

Model input parameters

(4) The model bottom boundary is fixed in all directions (ux =


uy = uz = 0).
(5) The ground surface of the model is free in all directions.
On average, each model consisted of approximately
10 000 elements (15-node wedge elements, with 6-noded
boundary elements) with an average size of 0.528 m, as depicted in Fig. 7. The mesh was refined in the vicinity of the
pile, that is, smaller dimension elements were generated, to
improve the accuracy of the model. Interfaces are composed
of 15-node interface elements of virtual thickness consisting of eight pairs of nodes, compatible with the 8-noded
quadrilateral side of a soil element. Along degenerated soil
elements, interface elements are composed of 6 node pairs,
compatible with the triangular side of the degenerated soil
element. Soilpile interfaces are considered smooth, in
that minimal relative displacements at the interface are allowed.
The numerical model was constructed to match the fullscale geometry of the pile in all regards excluding the
helical shape of the bearing plates, which were modeled as
circular discs rather than pitched plates. Additionally, the

The MohrCoulomb model was used to represent the soil


behaviour, for which cohesion and friction angle values
were obtained through limited triaxial test results and estimations based on SPT values. The soil parameters used in
the model were fine tuned to enhance the match between
the calculated and measured loaddeflection values for each
pile.
The process of establishing the models soil parameters
began by generating and loading a model for each pile at a
given site. Considering initially both total and effective
stress analyses, the high percentage of granular material
(which exhibits relatively rapid drainage) and the relatively
slow loading rate revealed that effective stress parameters
were absolutely the most representative for the analyses at
both sites. The loaddeflection relationships resulting from
the model simulations were plotted together with the load
deflection data obtained in the field for each pile. The model
soil parameters were then adjusted slightly until there was
sufficient agreement between loaddeflection curves from
the field and the model simulation results, for the majority
of the piles on the same site.
Figure 8 shows the loaddeflection curves from the FEM
and field load test for a pile tested in compression. Installed
to a depth of approximately 4.5 m, this pile rested in two
clayey silt layers. Because the measured and calculated
loaddeflection curves are in excellent agreement, as illustrated in Fig. 8, the soil properties used in this model are
#

2008 NRC Canada

1150

considered to be representative of the soil properties of the


two aforementioned soil layers for all other piles tested on
site 1. This process was repeated in cases where piles were
installed in different soil layers at the same site, until the entire soil stratigraphy was established and in good agreement
for all piles. The input parameters used in the model are
provided in Table 7 and Table 8. The high values of cohesion are indicative of overconsolidated soil.

Analysis of results
The piles tested in the field were loaded beyond the onset
of failure. In some cases, the failure detected in the FEM occurred at loads lower than those achieved in the field and as a
result, the numerical analysis could not be performed past the
onset of failure. This was primarily the result of the imposed
failure condition, in effect, when the applied load had to be
reduced in three successive calculation steps to reach equilibrium. Beyond this point, the soil body surrounding the
pile collapsed due to the imbalance between incremental
loads and soil shear strength; the result being large displacements for each successive, relatively small load increment. In
light of this, the loaddeflection curves generated by the FEM
were loaded to a level that matched the field loaddeflection
curves at least to the onset of rapid failure in all cases.

Load transfer mechanism


The interaction between the pile and surrounding soil was
examined from the output of the numerical models, and the
load transfer mechanism was established accordingly. Numerous analyses of the model output data were carried out
to characterize pile behaviour. This included inspection of
the state of stress within the soil in the vicinity of the pile
during the load. It also included examination of the strain
and displacement contours within the soil. Based on these
analyses, a conservative estimate of the shape of the model
failure surface was identified and an idealized failure mechanism was established. This was further used to define the
load transfer mechanism of the pile and to develop a model
to calculate the capacity of the helical piles.

Soil displacement
Monitoring soil displacements is a widely used technique
for estimating the failure surface of laboratory models. To
employ this approach in the numerical study, the total soil
displacements around the pile were noted at failure. The
contours of soil displacement are provided in increments of
20% of total pile displacement in Figs. 9a and 9b for piles
tested in compression and tension, respectively. These two
piles were found to behave in a manner most characteristic
of other piles under similar loading. The strain level beyond
the final shown contours was negligible (3 < 1  104).
Inspecting Fig. 9, the majority of the soil displacement
occurs within a radial distance of 1.5 helical diameters (of
largest helix) from the centre of the pile for both the tension
and compression cases. Additionally, the displacement contours follow immediately outside the annulus region enclosing the tapered helical profile extending from the top large
helix to the bottom small helix.
For the compressive loading case (Fig. 9a), the radial ex-

Can. Geotech. J. Vol. 45, 2008

tent of displacement contours are greatest at the level of the


top helix and are more pronounced for shallower piles.
Above the top helix, the displacement follows a conical
shape, reaching a maximum height of greater than one diameter of the uppermost helix. This region was shown to represent soil that is experiencing tensile stresses due to the
downward movement of the pile. A deep region of displacement below the bottom helix was present for all compression piles extending to a depth of greater than two
diameters of the lowest helix below the pile, which was
likely due to bearing of the bottom helix on the soil below.
The displacement contours for all compression piles suggest
that the failure surface follows a tapered cylinder shape that
roughly matches the interhelical profile. The load transfer
mechanism for compressive loading may therefore simplify
to a tapered pile-geometry extending between the top and
bottom helices with the bottom helix essentially representing
the pile toe. The toe resistance in this case may contribute
a significant portion of the pile capacity as the toe bears on
minimally disturbed soil.
For the case of uplift loading (Fig. 9b), all piles exhibit a
large region of soil displacement that extends above the top
helix to a height of greater than two diameters of the uppermost helix. This indicates significant bearing of the top helix
on the above soil, as suggested in the literature. Similar to
the compressive loading case, the greatest radial extent of
the displacement contours was near the top helix. This region was somewhat dissimilar however, to the rupture surface proposed by Ilamparuthi et al. (2002), which projects
outward from the top helix at an angle of 0.8. Based on
the results of the present model, the breadth of the displacement region was slightly greater than the diameter of the top
helix and tapered inward, rather than outward. However, it
should be noted that the investigation of Ilamparuthi et al.
(2002) was carried out on piles with lower embedment than
the piles modeled here.
Below the bottom helix, the displacement contours extended to a depth approximately equal to its diameter, enclosing a region of soil under tensile stresses. Therefore, the
soil displacements for the uplift case suggest a tapered failure surface that follows the interhelical profile combined
with substantial bearing of the top helix on the above soil.

State of stress
The state of stress within the soil around the pile was
noted at the failure load. The analysis of these stresses was
helpful in characterizing the soil zone where the stresses in
the soil approached the strength of the soil and failure was
likely to occur. It is therefore most convenient to assess the
state of stress in the soil in terms of the relative shear stress.
The relative shear stress, trel, is defined as
2

trel

t
tmax

where t* is the maximum value of shear stress (i.e., the radius of the Mohr stress circle). The parameter tmax is the
maximum value of shear stress for the case where the Mohrs
circle expanded to touch the Coulomb failure envelope, keeping the intermediate principal stress constant. Thus a relative
shear stress value of trel = 1 is indicative of soil failure.
#

2008 NRC Canada

Livneh and El Naggar

1151

Fig. 8. Measured and computed loaddeflection values for a compression pile in clayey silt.

Table 7. FEM soil properties at site 1.


Layer
Stiff brown sandy clayey silt
Very stiff brown clayey silt (W.T. depth)
Stiff grey clayey silt
Very stiff grey sandy clayey silt
Dense grey silt

Depth
(m)
2.4
4.1
5.8
7.3
>7.3

g
(kN/m3)
17.3
17.5
16.5
15
17

gsat
(kN/m3)
17.3
18.5
18.5
18
19

c
(kPa)
10
21
9
20
19


(8)
28
27
23
30
34

E
(kN/m2)
60 000
85 000
100 000
400 000
65 000

Note: W.T., water table; g, unit weight of the soil; gsat, saturated weight of the soil; c, cohesion of the soil; f, internal friction angle of the soil; E, Youngs modulus.

Table 8. FEM soil properties at site 2.


Layer
Stiff brown sandy clayey silt
Very stiff grey clayey silt (W.T. depth)
Dense grey fine sand, trace of silt

Depth
(m)
2.6
5.2
>5.2

g
(kN/m3)
18.5
17
18

gsat
(kN/m3)
18.5
19
20

c (kPa)
22
23
6


(8)
33
27
38

E
(kN/m2)
40 000
300 000
100 000

Note: W.T., water table; g, unit weight of the soil; gsat, saturated weight of the soil; c, cohesion of the soil; f, internal friction angle of the soil; E, Youngs modulus.

Examining the relative shear stress of piles tested in compression reveals similarities to the soil displacement contours (Fig. 9a), however, the region of soil approaching
failure tends to be much greater than that within a high
concentration of soil displacement contours. Figure 10a
shows a schematic of the full range of extent of the soil regions approaching shear failure (i.e., trel & 1) for all compression piles. Although the regions where the relative shear
stress approaches soil strength overestimates the size of the
failure region, they showed a consistent similarity to the
shape of the soil displacement geometry and served to confirm the general geometry of the failure zone that was used
to develop a design methodology.
Similar to the compression case, Fig. 10b shows a schematic of the range of relative shear stress failure extents for

all tension piles. In a similar manner, these extents resembled and confirmed the displacement contour failure geometry, with only a slightly larger volume of soil considered
to be approaching failure.

Design method for compressive loading


applications
Based on both measured and simulated behaviour of the
helical piles considered in this investigation, a method for
predicting the axial compressive capacity of the pile is presented.

Development of a failure surface


The regions of highest density of displacement contours
#

2008 NRC Canada

1152
Fig. 9. (a) Soil displacement contours for a compression pile as a
percentage of total pile displacement at failure. (b) Soil displacement contours for a tension pile as a percentage of total pile displacement at failure.

Can. Geotech. J. Vol. 45, 2008

Calculation of pile capacity


Under compression, it was shown that loads are transferred to the soil through two mechanisms: (i) shear resistance derived from a cylindrical failure surface along the
interhelical soil profile, and (ii) bearing of the bottom helix
on the soil below. A model is proposed for evaluating the
contribution of each of these load transfer mechanisms to
the pile capacity.
El Naggar and Sakr (2000) developed a technique for predicting the axial compressive capacity of tapered piles installed in sand. In this model, a taper coefficient, Kt, was
introduced to account for the taper effect on the skin friction, Qs, along the pile shaft, such that
3

Qstapered

Zl

Kt Ks v0 tanpdz

where Kt is the taper coefficient; Ks is the coefficient of lateral earth pressure; v0 is the effective overburden pressure;
d is the pilesoil interface friction angle (in this case d = ;
soilsoil interface); p is the pile perimeter; and l is the
length of the pile shaft.
The value of Kt is computed based on several factors, such
as the ratio, Sr, of pile settlement to diameter (= Up/D),
where pile settlement, Up, is evaluated using eq. [1], and Kt
is given by (El Naggar and Sakr 2000)
4

(i.e., greatest relative soil displacements) characterized the


pile failure surface. Considering the smallest extents of the
failure region to be representative ensures a consistent and
conservative solution, which is vital for design and application. Therefore, the following observations established the
geometry of the failure region for calculation of the compressive axial capacity of the system:
(1) Between the lowermost two helices, the failure surface is
observed to closely follow the profile of the helices and
is considered as such for design.
(2) At the level of the top helix, the failure surface is found
to extend a minimum radial distance of 30% of its diameter away from the helix. This value is used in the calculations.
(3) The extent of the failure surface above and below the
top and bottom helices was found to be highly variable,
and it is recommended that it be ignored for design purposes. Above the top helix, the soil was found to be in
tension, which will not contribute significantly to pile
capacity, while soil displacements below the bottom helix are attributed to the bearing resistance of this helix.
(4) The theoretical toe bearing of the bottom helix is used in
conjunction with the estimated shearing resistance between the helices to estimate pile capacity.
These observations are summarized in Fig. 11. The taper
angle, a, of the helices is approximately 2.128. The taper angle between the uppermost two helices is approximately
5.338 when considering the effect of the broad region of displaced soil contiguous to the top helix.

Kt A o

Bo
Sr
sv

where Ao and Bo depend on d, Ks, the elastic modulus of the


soil, G, and the taper angle, a, given by
5

Ao

tan cot
1 2tantan 

Bo

4Gtantan cot
1 2tantan Ks

In eq. [5], G is the shear modulus of the elastic soil; and


z lnr1 =rm , in which rm is the average pile radius, and r1
is the radius at which the shear stress becomes negligible
and is taken to be equal to 2.5/(1 n), where n is Poissons
ratio of the soil.
The contribution of toe bearing from the bottom helix, Qp,
was estimated using the approach outlined by Vesic (1963)
for drilled shafts. The bearing capacity factors associated
with drilled shafts were more representative of helical piles
and the influence of their installation disturbance on surrounding soil. Therefore, the toe bearing capacity of the bottom helix was estimated by
6

Qp Ap cNc q0 Nq

Combining eqs. [3] and [6], and considering two tapered


regions of cylindrical shearing resistance, the compressive
capacity of the pile is estimated as
7

Qu Qs5:33 Qs2:12 Qpbottom helix

The values of the pile capacity calculated using eq. [7]


#

2008 NRC Canada

Livneh and El Naggar

1153

Fig. 10. (a) Schematic covering the minimum extents for soil surrounding compression piles that is approaching relative shear stress failure
(i.e., trel = 1). (b) Schematic covering the minimum extents for soil surrounding tension piles that is approaching relative shear stress failure
(i.e., trel = 1). D, diameter of the top helix; Qu, axial capacity of the pile.

Fig. 11. Schematic of pile failure components for compressive


loading application. a, taper angle; D, diameter of the top helix; Qu,
pile capacity (under compression); Qp, toe bearing capacity of the
pile; Qs(tapered), skin friction capacity of the pile along the tapered
failure surface.

are summarized in Table 9 and are compared with measured


field values determined from the field load tests. The calculated capacities are found to be in good agreement with the
measured values to within 12% in all cases. The results indicate that this method is suitable for design, as it consistently yields a conservative estimate of pile capacity. The

design capacity of the pile can be obtained from the ultimate


capacity using an appropriate factor of safety.

Tension piles
The mode of failure is slightly different in the uplift case
as the effect of the tapered shaft does not improve pile capacity the way it does for the compression case (El Naggar
and Wei 1999). Numerous existing techniques were examined to predict the tensile capacity of the helical piles tested;
however, the method proposed by Mitsch and Clemence
(1985) for deep piles was found to be the most consistent,
to within approximately 20% of measured values for all
piles (Table 10). This method predicts the uplift capacity of
the pile by considering the algebraic sum of the bearing capacity of the uppermost helix and the frictional capacity developed at the interface of the interhelical soil using the
average helical diameter.
Pile spacing
To ensure that the design capacity of each helical pile is
fully mobilized, it is important to avoid interaction between
adjacent piles through the soil. It was noted from the numerical models that the soil outside the final displacement contour (i.e., 20% of total pile displacement) experienced very
low strain values; less than 3 = 1  104 and was less than
3 = 1  105 in many cases, that is, negligible for practical
purposes. The maximum radial extent of the influence zone
for a pile under compression was observed to reach a distance equal to two uppermost helical diameters (2D) from
the centerline of the pile and two and a half uppermost helical diameters (2.5D) for piles under tensile loads. Therefore,
for piles to mobilize their capacity fully and avoid interference between the influence zones of two adjacent piles,
centre-to-centre spacing should be a minimum distance of
#

2008 NRC Canada

1154

Can. Geotech. J. Vol. 45, 2008

Table 9. Comparison of measured and calculated bearing capacities of tested compression piles.

Compression piles
Pile 1, dense silt
Pile 2, dense silt
Pile 4, clayey silt
Pile 5, clayey silt
Pile 6, clayey silt
Pile 13, sand
Pile 14, sand

Measured
capacity,
Qu (kN)
465
496
356
327
328
830
728

Calculated
capacity,
Qt (kN)
455
458
329
319
310
738
708

Capacity
comparison
Qt/Qu (%)
97.75
92.43
92.31
97.57
94.51
88.95
97.25

(2)

(3)

Table 10. Comparison of measured and calculated bearing capacities of tested tension piles.

Tension piles
Pille 7, clayey silt
Pile 8(7 ft.), clayey silt
Pile 9, clayey silt
Pille 10(7 ft.), clayey silt
Pile 16, clayey silt
Pile 20, clayey silt
Pille 15, dense silt
Pile 17, dense silt
Pile 18*, dense silt
Pille 12, sand
Pile 19, sand

Measured
capacity,
Qu (kN)
226
285
181
305
250
356
381
256
178
378
406

Model
capacity,
Qt (kN)
175
260
170
285
235
226
350
232
160
322
356

Capacity
comparison
Qm/Qu (%)
77.43
91.23
93.92
93.44
94.00
63.48
91.86
90.63
89.89
85.19
87.68

(4)

(5)

*This performance is potentially misleading given the severe damage to


its bottom helix.

4D for compression applications and 5D for tensile applications.

Model uncertainties
The process of discretizing any continuous medium, such
as soil in this case, with a finite number of elements, will
contain some inherent approximations and inaccuracies. Additionally, the installation effects on the soil, which occurred
in the field, cannot be accurately modeled by the software
due to the absence of input data for the disturbed soil parameters. The accuracy of the analysis is also affected by
ignoring the increase of stiffness with depth, thus failing to
include both stress-dependency and stress-path dependency
of stiffness or anisotropic stiffness. However, the stress state
at failure is generally well described using the Mohr
Coulomb failure criterion with effective  and c parameters.

(6)

(7)

Conclusions
A comprehensive investigation was conducted into the axial performance of square-shaft helical piles. The findings of
the numerical model are in good agreement with the fullscale load test results, yielding the following noteworthy
conclusions:
(1) The loaddeflection curves of the piles tested displayed
typical trends, namely an initial linear segment, followed

by a highly nonlinear segment, and finally a near-linear,


rapid-failure segment. This was particularly consistent
and relevant for piles tested in compression, confirming
the suitability of helical piles for axial compressive loading applications.
A failure criterion was proposed to predict the ultimate
load for the piles tested. For cases where plunging failure did not occur promptly, the ultimate load is defined
as the load associated with deflection equal to 8% of the
diameter of the largesthelix plus the elastic settlement
PL
of the pile AE
0:08D .
The pile capacity was found to be proportional to the installation torque. Therefore, the empirical torque correlation coefficient Kt can be used to predict pile capacity.
In compression the value of Kt ranged between 35 and
42 m1 in dense and clayey silt, while in sand the value
was greater than 60 m1. In tension, the torque correlation coefficient was found to be 21.3 Kt 36.3 m1,
such that the upper range of values corresponded to piles
tested in dense silt and sand and the lower range was for
piles tested in clayey silt. Thus, the overall greater capacities of piles tested in compression translated into
greater torque correlation factors
The load transfer mechanism for all piles tested was
found to be predominantly through a tapered cylindrical
shear failure surface and bearing of the lead helix in
the direction of loading.
A method for estimating the axial compressive capacity
of helical piles was developed as the sum of the end
bearing and cylindrical shearing capacities of the pile.
The resistance derived along the cylindrical shearing surface was subdivided into two regions, each with a different taper angel, a, and was computed based on the
theory developed by El Naggar and Sakr (2000) for tapered piles. The first region has a taper angle of a =
5.338, extending from the top helix to the middle helix,
with an upper diameter 30% greater than that of the top
helix and a lower diameter equal to the diameter of the
middle helix. The second region has a taper angle of a =
2.128 extending from the middle helix to the bottom helix with an upper and lower diameter equal to that of
each respective helix. Finally, the end bearing of the bottom helix was computed using the bearing capacity factors presented by Vesic (1963) for drilled shafts.
Estimates of pile capacity using this technique were
shown to be conservative and accurate within 12% of
measured values in all cases.
The uplift capacity was developed in a slightly different
manner and was most consistently computed by the
method outlined by Mitsch and Clemence (1985).
To avoid overlapping between the influence zones of adjacent helical piles, a minimum spacing of 4D and 5D
was proposed for piles loaded in compression and tension, respectively.

Acknowledgements
The donation of the helical piles by Hubbell Power Systems Inc. is greatly appreciated. The field installation of the
helical piles by EBS Engineering and Construction is also
greatly appreciated.
#

2008 NRC Canada

Livneh and El Naggar

References
ASTM. 1994. Standard test method for individual piles under static
axial compressive load, D114381 (reapproved 1994). ASTM
International, West Conshohocken, Pa.
ASTM. 1995. Standard test method for individual piles under static
axial tensile load, D368990 (reapproved 1995). ASTM International, West Conshohocken, Pa.
Balla, A. 1961. The resistance to breaking-out of mushroom foundations for pylons. In Proceedings of the 5th International Conference of Soil Mechanics and Foundation Engineering, Paris,
France, Vol. 1, pp. 569576.
BSI. 1986. British standard code of practice for foundations. BS
8004. British Standards Institution (BSI). London.
Butler, H.D., and Hoy, H.E. 1977. The Texas quick-load method
for foundation load testing Users manual. Rep. No. FHWAIP-778. Texas State Department of Highways and Public
Transportation, Austin, Tex.
Clemence, S.P.N., and Veesaert, C.J. 1977. Dynamic pullout resistance of anchors in sand. In Proceedings of the International
Symposium On Soil-Structure Interaction, Roorkee, India.
pp. 389397.
Davisson, M.T. 1972. High capacity piles. In Proceedings of Lecture Series of Innovations in Foundation Construction, ASCE, Illinois section, Chicago. pp. 81112.
Downs, D.I., and Chieurzzi, R. 1966. Transmission tower foundations. Journal of the Power Division, 92(P02): 91114.
El Naggar, M.H., and Sakr, M. 2000. Evaluation of axial performance of tapered piles from centrifuge tests. Canadian Geotechnical Journal, 37: 12951308. doi:10.1139/cgj-37-6-1295.
El Naggar, M.H., and Wei, J.Q. 1999. Axial capacity of tapered
piles established from model tests. Canadian Geotechnical Journal, 36: 11851194. doi:10.1139/cgj-36-6-1185.
FDOT. 1999. Standard specifications for road and bridge construction. Florida Department of Transportation (FDOT). Tallahassee,
Fla.
Ghaly, A., and Hanna, A. 1991. Experimental and theoretical studies on installation torque of screw anchors. Canadian Geotechnical Journal, 28: 353364. doi:10.1139/t91-046.
Ghaly, A.M., Hanna, A.M., and Hanna, M.S. 1991. Uplift behavior
of screw anchors in sand: I. Dry sand. Journal of Geotechnical
Engineering,
117:
773793.
doi:10.1061/(ASCE)07339410(1991)117:5(773).
Hoyt, R.M., and Clemence, S.P. 1989. Uplift capacity of helical anchors in soil. In Proceedings of the 12th International Conference on Soil Mechanics and Foundation Engineering (ISSMFE),
Rio de Janeiro, Brazil. Vol. 2. pp. 10191022.
Hubbell Power Systems Inc. 2003. Industry standards based on
CHANCE multi-helix anchor specs. Bulletin 040301. Centralia,
Mo.
Ilamparuthi, K., Dickin, E.A., and Muthukrisnaiah, K. 2002. Experimental investigation of the uplift behaviour of circular plate
anchors embedded in sand. Canadian Geotechnical Journal,
39(3): 648664. doi:10.1139/t02-005.
ISSMFE. 1985. Axial pile loading test Part I: Static loading.
Geotechnical Testing Journal, ASTM, 8(2): 7980
Ireland, H.O. 1963. Discussion, uplift resistance of transmission
tower footings. Journal of the Power Division, 89: 115118.

1155
Majer, J. 1955. Zur berechnung von zugfundamenten. Osterreichische Bauzeitgschrift, 10: 8590. [In German.]
Meyerhof, G.G., and Adams, J.I. 1968. The ultimate uplift capacity
of foundations. Canadian Geotechnical Journal, 5(4): 225244.
doi:10.1139/t68-024.
Mitsch, M.P., and Clemence, S.P. 1985. The uplift capacity of helix anchors in sand. In Proceedings of the Uplift Behavior of
Anchor Foundations in Soil, Detroit, Mich., 24 October 1985
Edited by S.P. Clemence. American Society of Civil Engineers.
pp. 2647.
Mors, H. 1959. The behavior of mast foundations subjected to tensile forces. Bautechnik, 36: 367378.
Murray, E.J., and Geddes, J.D. 1987. Uplift of anchor plates in
sand. Journal of Geotechnical Engineering, 113: 202215.
Narasimha Rao, S., and Prasad, Y.V.S.N. 1991. Estimation of uplift
capacity of helical anchors in clays. Journal of Geotechnical Engineering, 199: 352357.
Narasimha Rao, S., Prasad, Y.V.S.N., Shetty, M.D., and Joshi, V.V.
1989. Uplift capacity of screw anchors. Geotechnical Engineering, 20: 139159.
Narasimha Rao, S., Prasad, Y.V.S.N., and Veeresh, C. 1993. Behaviour of embedded model screw anchors in soft clays. Geotechnique, 43: 605614.
Prakash, S., and Sharma, H.D. 1990. Pile foundations in engineering practice. John Wiley & Sons, New York. p. 824.
Reese, L.C., and ONeill, M.W. 1988. Drilled shafts: construction
procedures and design methods, FHWA-HI-88042. Federal
Highway Administration, McLean, Va.
Rowe, R.K., and Davis, E.H. 1982. The behaviour of anchor plates
in sand. Geotechnique, 32: 2541.
SAA. 1995. Piling design and installation. Australian standard AS
2159. Standards Association of Australia (SAA), Homebush,
NSW.
Saeedy, H.S. 1987. Stability of circular vertical anchors. Canadian
Geotechnical Journal, 24: 452456. doi:10.1139/t87-056.
Sutherland, H.B. 1965. Model studies for shaft raising through cohesionless soils. In Proceedings of the 6th International Conference on Soil Mechanics and Foundation Engineering, Montreal,
Que. Vol. 2, pp. 410413.
Sutherland, H.B., Finlay, T.W., and Fadl, M.O. 1982. Uplift capacity of embedded anchors in sand. In Proceedings of the 3rd International Conference on the Behavior of Offshore Structures,
Massachusetts Institute of Technology, Cambridge, Mass.
pp. 451463.
Vesic, A.S. 1963. Bearing capacity of deep foundations in sand.
Highway Research Record, No. 39, Highway Research Board,
National Academy of Science, Washington, D.C., pp. 112153.
Vesic, A.S. 1971. Breakout resistance of objects embedded in
ocean bottom. Journal of the Soil Mechanics and Foundation
Engineering Division, ASCE, 97: 11831205.
Weizhi, S., and Fragaszy, R.J. 1988. Uplift testing of model anchors. Journal of Geotechnical Engineering, 114: 961983.
Zhang, L.M., Li, D.Q., and Wilson, H.T. 2005. Reliability of bored
pile foundations considering bias in failure criteria. Canadian
Geotechnical Journal, 42: 10861093. doi:10.1139/t05-044.

2008 NRC Canada

Potrebbero piacerti anche