Axial Compressive Response of Large-Capacity Helical and Driven Steel Piles in Cohesive Soil

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/272433345

Axial compressive response of large-capacity helical and driven steel piles in


cohesive soil

Article  in  Canadian Geotechnical Journal · February 2015


DOI: 10.1139/cgj-2012-0331

CITATIONS READS
32 1,249

2 authors:

Mohamed Elkasabgy M.Hesahm El Naggar


Benha University The University of Western Ontarioprofessor and Associate Dean
14 PUBLICATIONS   204 CITATIONS    395 PUBLICATIONS   6,058 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Geo-Structural Nonlinear Analysis Program (GSNAP) View project

Stone Column Reinforced Soils View project

All content following this page was uploaded by M.Hesahm El Naggar on 09 March 2015.

The user has requested enhancement of the downloaded file.


224

ARTICLE
Axial compressive response of large-capacity helical and driven
steel piles in cohesive soil
Mohamed Elkasabgy and M. Hesham El Naggar
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by University of Western Ontario on 02/17/15

Abstract: The axial compression performance of large-capacity helical piles is of significant interest because they can offer an
efficient alternative to conventional piling systems in many applications such as in oil processing facilities, transmission towers,
and industrial buildings. This paper presents the results of seven full-scale axial compression load tests conducted on 6.0 and
9.0 m large-capacity helical piles and a 6.0 m driven steel pile. The results are considered essential to qualify and quantify the
performance characteristics of large-capacity helical piles in cohesive soils. The test piles were close-ended steel shafts with an
outer diameter of 324 mm. The test helical piles were either single or double helix, with a helix diameter of 610 mm and
interhelix spacing that varied between 1.5 and 4.5 times the helix diameter. The subsurface soil properties at the test site were
determined using field and laboratory testing methods. The 6.0 m piles were tested 2 weeks after installation, while the 9.0 m
piles were tested 9 months after installation. The load–settlement curves were presented to better understand the behaviour of
test piles. An ultimate capacity criterion was proposed to estimate the ultimate load of large-capacity helical piles. The test
helical piles developed ultimate resistances up to 1.2–1.8 times that of the driven pile. The load-transfer mechanisms of large-
capacity helical piles were studied, and it was found that soil disturbance during pile installation had a significant effect on the
pile failure mechanism regardless the value of the interhelix spacing to helix diameter ratio. The mobilized soil strength
parameters were back-calculated and compared with the estimated intact soil strength parameters.

Key words: helical piles, driven piles, full-scale testing, compression, cohesive soils, load transfer, cone penetration test.
For personal use only.

Résumé : La performance en compression axiale de pieux hélicoïdaux à grande capacité comporte un intérêt significatif puisque
ceux-ci peuvent offrir une alternative efficace aux systèmes de pieux conventionnels dans plusieurs applications, telles que les
installations de traitement du bitume, les tours de transmission et les édifices industriels. Cet article présente les résultats de
sept essais de compression axiale à l’échelle réelle réalisés sur des pieux hélicoïdaux à grande capacité de 6,0 et 9,0 m et sur un
pieu foncé en acier de 6,0 m. Les résultats sont considérés comme essentiels pour qualifier et quantifier les caractéristiques de
la performance des pieux hélicoïdaux à grande capacité placés dans des sols cohésifs. Les pieux d’essai avaient des arbres d’acier
fermés aux extrémités avec un diamètre de 324 mm. Les pieux d’essai hélicoïdaux avaient une hélice simple ou double, avec un
diamètre d’hélice de 610 mm et un espacement inter-hélice qui variaient entre 1,5 et 4,5 fois le diamètre de l’hélice. Les propriétés
du sol au site d’essai ont été déterminées à l’aide de méthodes expérimentales en laboratoire et sur le terrain. Les pieux de
6,0 m ont été testés deux semaines après leur installation, tandis que les pieux de 9,0 m ont été testés neuf mois après
l’installation. Les courbes de charge–tassement sont présentées pour offrir une meilleure compréhension du comportement des
pieux d’essai. Un critère de capacité ultime est proposé pour estimer la charge ultime de pieux hélicoïdaux à grande capacité. Les
pieux hélicoïdaux testés ont développé des résistances ultimes jusqu’à 1,2 à 1,8 fois celle du pieu foncé. Les mécanismes de
transfert de charge des pieux hélicoïdaux à grande capacité ont été étudiés, et il a été démontré que le remaniement du sol
durant l’installation du pieu a un effet significatif sur le mécanisme de rupture du pieu peu importe la valeur du ratio de
l’espacement inter-hélice sur le diamètre de l’hélice. Les paramètres de résistance mobilisée du sol sont rétro-calculés et
comparés aux paramètres de résistance estimés pour un sol intact. [Traduit par la Rédaction]

Mots-clés : pieux hélicoïdaux, pieux foncés, essai à l’échelle réelle, compression, sols cohésifs, transfert de charge, essai de
pénétration du cône.

Introduction Helical piles are recognized as a viable foundation option in


A helical pile consists of one or more helical-shaped circular several engineering applications because of their ease of installa-
plate(s), or helix, affixed to a steel central shaft as illustrated in tion, especially in limited access areas with low levels of noise and
Fig. 1. Based on geometrical and design considerations, two helical vibration, and cost effectiveness. In addition, helical piles allow
pile categories can be practically identified, namely small-capacity immediate loading upon installation and can be installed through
helical piles and large-capacity helical piles. A large-capacity heli- groundwater without casing, unaffected by caving soils (Bobbitt
cal pile is formed of a steel pipe of diameter >178 mm (7 in.) with and Clemence 1987; Perko 2009).
a hollow circular cross section. Helical piles can be installed with Analytical methods have been developed to calculate the
shaft diameters up to 914 mm (36 in.) and with axial compression helical pile capacity under compression or uplift loading (Mitsch
capacities of up to 4500 kN, depending on pile geometry and and Clemence 1985; Mooney et al. 1985; Narasimha Rao et al. 1991,
subsurface conditions. 1993). Several design methods were established based on the

Received 6 September 2012. Accepted 3 July 2014.


M. Elkasabgy. AMEC Environment and Infrastructure, 140 Quarry Park Blvd. SE, Calgary, AB T2C 3G3, Canada.
M.H. El Naggar. Geotechnical Research Centre, Faculty of Engineering, The University of Western Ontario, London, ON N6A 5B8, Canada.
Corresponding Author: M. Hesham El Naggar (e-mail: helnaggar@eng.uwo.ca).

Can. Geotech. J. 52: 224–243 (2015) dx.doi.org/10.1139/cgj-2012-0331 Published at www.nrcresearchpress.com/cgj on 17 July 2014.
Elkasabgy and El Naggar 225

Fig. 1. Failure modes: (a) cylindrical shear method; (b) individual plate bearing method. D, helix diameter; d, pile shaft diameter; H, embedment
depth of upper helix; Heff, pile shaft effective length; Lc, distance between top and bottom helices; S, interhelix spacing.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by University of Western Ontario on 02/17/15
For personal use only.

individual plate anchor theories (e.g., Adams and Hayes 1967; sand and in clay and proposed a method for their design based on
Meyerhof and Adams 1968), while some methods are similar to the results of the calibrated numerical model.
those proposed for underreamed cast-in-place piles (Narasimha The installation of helical piles in cohesive soils can cause re-
Rao et al. 1993). The helical plates increase the helical pile capacity moulding of the adjacent cohesive soil within the zone affected by
significantly as do the underreams in the underreamed piles. the pile helices. This remoulding can introduce fissures within the
Experimental studies demonstrated that the behaviour of helical pile cylindrical failure surface and surface heaving (Mooney et al.
piles under compression and tension loads is similar (Narasimha 1985; Bradka 1997; Zhang 1999). Thus, the shear strength of cohe-
Rao et al. 1991; Zhang 1999). Thus, the design methods are consid- sive soil decreases along the affected zone. In cohesionless soils,
ered to be interchangeable. screwing the pile induces stress changes in the soil due to the
Tappenden (2007) investigated the accuracy of the established accompanied soil disturbance. The disturbed zone may become
design methods with which the axial capacity of helical piles may weak, which causes a cylindrical failure surface around the pile
be predicted using the results of 29 full-scale axial load tests in (Vesic 1971). Weech (2002) studied the soil disturbance caused by
tension and compression for piles installed in western Canada. the installation of full-scale small-diameter helical piles in sensi-
Lutenegger (2009) indicated that there was no distinct transition tive fine-grained soils and the mobilization of axial pile capacity at
from cylindrical shear to individual bearing behaviour based on different times after installation. Vyazmensky (2005) developed a
the results of a field study that was conducted to investigate the numerical formulation that incorporated a simplified technique
behaviour of multi-helix anchors in clay. Sakr (2009) investigated that allowed reasonable prediction of stresses and pore pressure
the axial capacity of large-diameter helical piles installed in oil variation during and after helical pile installation based on the
sands. results of the field testing performed by Weech (2002).
Several numerical studies were conducted to investigate the Clemence and Pepe (1984) and Mitsch and Clemence (1985) per-
axial compressive and uplift characteristics of helical piles and formed field and laboratory tests and reported that during helical
anchors. Merifield et al. (2006) evaluated the effect of the shape of anchor installation, the sand in contact with the helices was dis-
the anchor plate (i.e., square, circular, or rectangular) on the axial placed laterally, which in turn densified the surrounding sand
capacity of horizontal anchors in undrained clay through three- and increased the lateral stress. However, depending on the den-
dimensional (3D) finite element modelling. Livneh and El Naggar sity of sand, stress level, and shear strain level, dilation may or
(2008) investigated the load-transfer mechanism of the square- may not take place.
shaft slender helical piles using 3D finite element models. Kurian Based on observed failure modes of helical piles installed in
and Shah (2009) developed 3D and axisymmetric finite element cohesive and cohesionless soils, there are two common methods
models to study the axial and lateral capacities of helical piles in to account for the contribution of the helices to the pile capacity,
drained cohesive soils. Elsherbiny and El Naggar (2013) investi- namely the cylindrical shear method and the individual plate
gated the axial compressive behaviour of smaller helical piles in bearing method. In the cylindrical shear method (Mooney et al.

Published by NRC Research Press


226 Can. Geotech. J. Vol. 52, 2015

Fig. 2. Plan view of piles, seismic cone penetration tests, and borehole. SPT, standard penetration test.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by University of Western Ontario on 02/17/15
For personal use only.

1985), a cylindrical failure surface is assumed between the top upper helix, H, minus the helix diameter), and ␣ is the adhesion
and bottom helices (Fig. 1a). The ultimate compressive capacity factor. Zhang (1999) stated that the method provided better capac-
of the pile is the sum of the bearing resistance of the bottom helix, ity prediction for multi-helix piles with interhelix spacing (S) to
the frictional resistance along the cylindrical failure surface, and diameter (D) ratio of S/D ≤ 3.0.
the frictional resistance along the pile shaft. In the individual For cohesionless soils, the formulation proposed by Mitsch and
plate bearing method (Meyerhof and Adams 1968), the pile com- Clemence (1985) can be used to estimate the compression capacity:
pressive capacity is calculated assuming an individual bearing
failure occurs below each helix plate (Fig. 1b). Thus, the ultimate
compressive capacity of the pile is the sum of the individual bear-
(2) Q ult ⫽ (Ah ⫹ At)␥ HnNq ⫹ 共 ␲2 DL 兲K tan␾ ⫹ ␲2 dH
2
c s
 2 
eff␥ Ks tan␾ 

ing resistance of all helices and the frictional resistance along the
pile shaft. where Hn is the depth to the bottom helix, ␥= is the effective soil
unit weight, Nq is the bearing capacity factor, Ks is the coefficient
Objectives and scope of work of lateral earth pressure, and ␾  is the effective angle of internal
friction. Zhang (1999) reported that the cylindrical shear method
The main objective of this study is to investigate the perfor-
provides a better prediction of the compression capacity of piles
mance characteristics of single large-capacity helical and driven
in sand with S/D ≤ 2.0.
piles installed in cohesive soils and subjected to axial compression
loading. The specific objectives of the testing program were to Individual plate bearing method
(i) determine the load-transfer mechanisms of helical piles, For cohesive soils, the individual plate bearing method can be
(ii) define the appropriate failure criterion for large-capacity heli- adopted for piles with S/D > 3.0 (Zhang 1999). The ultimate capac-
cal piles, and (iii) evaluate the axial compression capacities of ity of helical piles in cohesive soils is calculated as follows:
tested piles. The testing program included seven full-scale axial
compression load tests carried out on six large-capacity helical
piles and one driven steel pile.
(3) Q ult ⫽ AtCuNc ⫹ 兺 (A C N ) ⫹ (␲dH
h u c eff)␣Cu

Failure modes of helical piles For cohesionless soils, the individual plate bearing method can be
Cylindrical shear method used to calculate the ultimate capacity for the case of S/D > 2.0, i.e.,
The ultimate compression capacity, Q ult, of helical piles in co-
hesive soil may be given by (4) Q ult ⫽ At␥ HnNq ⫹ 兺 A ␥ H N ⫹ ␲2 dH
i⫽1⫺n
h

i q
2 
eff␥ Ks tan␾ 

(1) Q ult ⫽ (Ah ⫹ At)CuNc ⫹ (␲DLc)Cu ⫹ (␲dHeff)␣Cu


where Nq and Hi are the bearing capacity factor and embedment
where Ah is the projected area of the helix, At is the area of the pile depth of the upper helices, respectively.
tip, Cu is the undrained soil cohesion, Nc is the bearing capacity
factor, d is the pile shaft diameter, D is the helix diameter, Lc is the Pile configurations and distribution
distance between the top and bottom helices, Heff is the effective Six helical piles and one driven pile were manufactured and
length of the pile shaft (equal to the embedment depth of the installed. Four of the helical piles were 6.0 m long and two were

Published by NRC Research Press


Elkasabgy and El Naggar 227

Table 1. Geometrical properties of test and reaction piles.


No. of Interhelix Installation
Pile No. Type Length (m) helices spacing, S (m) H/D S/D torque, T (kN·m) KT (m−1)
Phase I (piles tested 2 weeks after installation)
SS11 Test (helical) 6.0 1 — 8.9 — 83.0 10.2
SD11 Test (helical) 6.0 2 0.9 7.5 1.5 115.2 11.6
SD21 Test (helical) 6.0 2 1.8 6.0 3.0 112.9 10.0
SD31 Test (helical) 6.0 2 2.7 4.5 4.5 92.2 11.3
DS Test (driven) 6.0 — — — — — —

Phase II (piles tested 9 months after installation)


Can. Geotech. J. Downloaded from www.nrcresearchpress.com by University of Western Ontario on 02/17/15

LS12 Test (helical) 9.0 1 — 13.9 — 94.9 20.5


LD12 Test (helical) 9.0 2 0.9 12.4 1.5 93.4 26.5
R Reaction (helical) 6.0 2 2.1 7.9 3.0 — —
Note: S, interhelix spacing; H/D, embedment ratio defined as the ratio of the embedment depth of the upper helix, H, to the helix
diameter, D; T, pile installation torque; KT, empirical capacity-to-torque factor.

9.0 m long. Two helical piles were fabricated as single-helix piles, brown and dense to very dense silty sand; layer 5, a 7.2 m thick
and the other four piles were double-helix piles with different layer of grey and overconsolidated very stiff to hard clay till of
interhelix spacing ratios, S/D = 1.5–4.5. The driven pile was 6.0 m medium to high plasticity with interbedded seams of layers of
long and had the same shaft diameter and wall thickness as the sandy silt, silty sand, and silt at a depth of 8.0 m; layer 6, interbed-
test helical piles. Figure 2 and Table 1 provide the distribution and ded layers of moderately to heavily overconsolidated sandy silt,
basic geometrical characteristics of the test and reaction piles silt, clayey silt, and silty clay. The pore pressure measurements
employed for compression load tests. The piles consisted of a and dissipation tests obtained from the cone tests indicated
central steel shaft of outer diameter 324 mm and wall thickness groundwater level at 1.2 m below the ground surface.
9.5 mm. The helix plate was 610 mm in diameter, with helix pitch A summary of the soil index properties, stress history parame-
of 152 mm (6 in.), and thickness of 19 mm. All test piles were ters, and strength properties obtained from laboratory tests and
close-ended with a flush closure steel plate. The pile shafts con- from correlations with SPT and cone penetration test (CPT) results
For personal use only.

form to the material and dimensional requirements of ASTM stan- are presented in Table 2.
dard A252-98 (ASTM 2007a), grade 2 steel. The tensile and yield
strengths of the piles’ material were 414 and 240 MPa, respec- Pile installation
tively, and the elastic modulus of the pile material was 210 GPa. Figure 5 shows a schematic diagram of the test piles. The helical
The test layout complied with the specifications of the ASTM piles were installed by applying a clockwise turning moment
standard D1143-07 (ASTM 2007b). The axial compression testing (torque) to the pile shaft, while a rate of penetration of one helix
program included seven full-scale load tests, which were con- plate pitch (152.4 mm) per revolution was considered to reduce
ducted in two phases: phase I and phase II (Table 1). Phase I was soil disturbance. The driven pile was installed using a mechanical
carried out on the 6.0 m helical piles (SS11, SD11, SD21, and SD31) drop hammer, weighing 19 kN, and dropped through a distance of
and the 6.0 m driven pile (DS), 2 weeks after pile installation, about 0.9 m onto the pile cushion. The piles protruded 0.6 m
when the soil around the piles’ shaft was disturbed because of the above the ground surface (i.e., unsupported length).
pile installation process. Phase II of the piles testing program was Generally, installation of helical piles causes disturbance to co-
carried out on the 9.0 m helical piles (LS12 and LD12), 9 months hesive soils. This disturbance and associated reduction in strength
after pile installation, which was deemed to be enough time for and stiffness are even more pronounced for structured cohesive
the disturbed soil to regain significant amount of its original stiff- soils. This type of soils exhibits higher degradation in strength
ness and strength. and stiffness when the natural microstructure is disturbed.
The disturbance can be attributed to the advancement of the
Subsurface conditions pile shaft into the ground and the rotation of the helical plates. In
The test site is located north of the town of Ponoka, in the the current study, the level of disturbance due to shaft penetra-
central Alberta region, Canada. The surficial deposits in the test tion is expected to be higher, as the piles were installed close-
site zone consist of Pleistocene Stagnation Moraine glacial till ended. Therefore, it can be assumed that the test helical piles
depositions (Shetsen 1990). Both in situ and laboratory tests were acted as displacement piles during the installation process in a
performed to characterize the index, stress history parameters, way that the site's structured cohesive soils were displaced and
and shear strength properties of the site soils. Laboratory tests remoulded around the pile tip. Furthermore, slight ground sur-
were conducted on disturbed and partially undisturbed samples face heave was monitored during the installation of the test piles.
collected from one borehole at an interval of 1.5 m. Standard In addition, the soil displacement and shearing phenomena
penetration tests (SPTs) were performed in the borehole to deter- would cause the soil structure to collapse and lead to a buildup of
mine the N value at different depths. Three seismic cone penetra- excess pore-water pressure that causes a reduction in effective
tion tests (SCPTs) were conducted within the site (see Fig. 2) to a stress. This temporary buildup of excess pore-water pressure
depth of 15 m. The SCPT results were used to determine the soil would cause the cohesive soil to experience significant reduction
layering as shown in Figs. 3 and 4. in its shear strength.
The subsurface conditions at the test site (Figs. 3 and 4) consist The contribution of the helical plates to the disturbance of the site
of the following: layer 1, a 1.0–1.5 m thick brownish and medium soil can be ascribed to different phenomena. The helix disturbs soil
dense sandy silt crust with some organic materials; layer 2, a through a spirally shaped cut with intervals equal to the pitch dis-
2.8–3.1 m thick layer of brownish and moderately overconsoli- tance, in which the soil traversed by the helix is sheared and dis-
dated stiff clay to silty clay and clayey silt of medium plasticity placed laterally and vertically. However, this level of disturbance
interbedded with seams of silt; layer 3, 0.6–1.3 m of interbedded may increase for layered subsurface conditions, where crowding is
layers of grey and overconsolidated very stiff silty clay, clayey silt, employed to advance the pile into harder parts of the soil profile.
and clay of medium plasticity; layer 4, a 0.3–1.5 m thick yellowish Furthermore, a sudden change in the applied installation torque due

Published by NRC Research Press


228 Can. Geotech. J. Vol. 52, 2015

Fig. 3. Measured CPT profile at SCPT-2 (1 bar = 100 kPa). W.L., groundwater level.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by University of Western Ontario on 02/17/15

Fig. 4. Schematic diagram for soil stratigraphy cross sections along (a) 6.0 m and (b) 9.0 m helical piles. ␥, bulk unit weight.
For personal use only.

to the change in strength and stiffness parameters between the soil ticles also prevented advancing the cone soundings SCPT-1 and
layers would cause alteration in the rate of the pile penetration, SCPT-2, which were relocated to the new locations shown in Fig. 2.
which can cause more soil disturbance especially for multi-helix The torque values required to install the double-helix piles
piles. SD11, SD21, and SD31 were ⬃11%–39% higher than the torque value
The final measured average torques at the end of helical pile required for the single-helix pile SS11. It can be seen that the
installation and total embedment depths are given in Table 1. It torque value required for the double-helix pile SD31 is smaller
should be noted that the distribution of the torque values along than those values recorded for the double-helix piles SD11 and
the full embedment depths of test piles and the driving record of
SD21 by ⬃19%–20% as a result of the interhelix spacing and soil
the driven pile were not available by the manufacturing company.
layering. In addition, the torque values required to install the
It can be noticed from Table 1 that torque values for piles SS11,
9.0 m single-helix and double-helix piles LS12 and LD12, tested in
SD11, SD21, and SD31, tested in phase I, varied between 83 and
115.2 kN·m. As explained later, the torque values were affected by phase II, were 94.9 and 93.4 kN·m, respectively.
the presence of large-size particles, which could be the possible
cause of having large torque values for the double-helix piles SD11 Pile instrumentation
and SD21, as some difficulties were experienced in installing those The helical piles were instrumented with 7–12 levels of strain
piles to the designated depth. The presence of the large-size par- gauges, depending on the pile length and number and spacing of

Published by NRC Research Press


Elkasabgy and El Naggar 229

Table 2. Summary of soil properties.


Property Layer 1 Layer 2 Layer 3 Layer 4 Layer 5 Layer 6
Soil type Silt crust
Clay to silty clay Silty clay, clayey silt, Silty sand Clay till Sandy silt, clayey silt,
and clayey silt and clay silt, and silty clay
Subsurface condition Medium dense Stiff Very stiff Dense to very dense Very stiff to hard Very stiff
Thickness (m) 1–1.5 2.8–3.1 0.6–1.3 0.3–1.5 7.2 1.7
␥ (kN/m3) 18.2 17.8 17.8 18.8 17.5 18.0
Wc (%) 16 17 17 18 17–28 —
WLL (%) — 30–39 30–39 — 32–47 —
IP (%) — 14–17 14–17 — 12–29 —
␯ 0.35 0.42 0.46 0.35 0.48 0.49
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by University of Western Ontario on 02/17/15

OCR 10 10 14 6 15 7
Cu (kN/m2) — 85 137 — 177 150
␾  (°) 31 — — 42 — —
Note: ␥, bulk unit weight; Wc, natural moisture content; WLL, liquid limit; IP, plasticity index; ␯, Poisson’s ratio (drained value for layers 1 and 4 and undrained value
for layers 2, 3, 5, and 6); OCR, overconsolidation ratio; Cu, undrained soil cohesion; ␾ , angle of internal friction. ␥, ␯, OCR, Cu, and ␾  were estimated from the results
of the SCPT and SPT tests.

helices (see Fig. 5). Half-bridge strain gauge circuits were affixed beam was placed on the two reaction piles and centered over the
on the inner wall of the piles at specified locations. For the 6.0 m test pile.
piles, the first level of gauges were installed on the outer surface A steel cap plate of nominal thickness of 40 mm was centered
of the piles, while the first, second, and third levels of gauges were and welded on the head of each test pile. Two plates with thick-
installed on the outer surface for the 9.0 m piles. Each level of nesses of 25 and 14 mm were placed on top of the cap plate. The
gauges encompassed four half-bridges allocated equidistantly hydraulic jack was placed on top of the steel cap, and the load cell
along the shaft circumference. The pile was cut into short seg- was situated between the jack and the main reaction beam using
ments, the gauges were installed at an approximate distance of steel plates of nominal thickness of 25 mm in between.
250 mm from the edge of each segment, and then gauges and
wires were covered with a fiber cloth and epoxy. The pile seg- Displacement measurements
For personal use only.

ments were then welded together to form the test pile while care The vertical pile movement was measured using four calibrated
was exercised to ensure ground flush and grind welds. The strain linear displacement transducers (LDTs) with accuracy of 0.01 mm
gauges and wires were protected by fiberglass insulation from the and maximum mechanical travel of 100 mm. The transducers
heat during the welding process. were mounted on two independently supported 102 mm box-
section steel reference beams. The lateral movement of the test
Test setup and procedures pile was monitored using two dial gauges of accuracy 0.25 mm
Schematic diagrams and photos of the load test setups are illus- (0.01 in.).
trated in Fig. 6 for helical piles and driven pile. The following
sections describe the setup loading system, the test reaction sys- Test procedures
tem, the displacement measurement devices, and test procedures. The axial compression load test was carried out following the
quick maintained load test method recommended by ASTM stan-
Loading system dard D1143-07 (ASTM 2007b). Each pile was loaded in increments of
A calibrated hydraulic jack, actuated with a hydraulic pump, 10%–15% of the anticipated ultimate load. Each increment was
with capacity of 1800 kN was used to apply the compression load, maintained for a period of 5 minutes, with displacement and
while a calibrated load cell of capacity 4500 kN was employed to strain readings recorded by the data acquisition system. The load
measure the load. For the double-helix pile LD12, two hydraulic increments were applied until continuous jacking was required to
jacks each of capacity 1800 kN were necessary to apply the com- maintain the test load, which was considered as failure of the test
pression load. A hydraulic pressure transducer, with capacity of pile. The pile was then unloaded by removing the test load in four
69 MPa, was attached to the hydraulic jack to measure the pres- equal decrements, while keeping the load constant for a time
sure applied at the pile head. A main beam and two secondary interval between 2.5 and 5 minutes. When reaching a zero applied
beams were utilized as the reaction frame for testing helical piles. load, enough time was allowed to monitor the pile head displace-
For the driven pile, the reaction frame comprised only one main ment readings to assess the rebound behaviour. Most tests were
beam. carried out up to a pile head settlement of ⬃60 mm, which corre-
sponded to a settlement of 10% of the helix diameter.
Reaction system
The reaction system for testing helical piles was composed of Load test results
four reaction helical piles, each of shaft diameter 324 mm, length
6.0 m, helix diameter 685 mm, and interhelix spacing of three Load–settlement curves
times the helix diameter. The reaction piles had at least 4.5 times This section presents the results of the axial compression
the capacity of the 6.0 m test piles and at least 1.6 times the load tests conducted on piles tested in phases I and II. The load–
capacity of the 9.0 m test piles. However, the reaction system for settlement curves of phase I are given in Figs. 7a and 7b for the
the test driven pile was composed of two helical piles, with a helical piles (SS11, SD11, SD21, and SD31) and the driven pile (DS),
capacity of at least 2.4 times the capacity of the driven pile. The respectively. Figure 7c presents the load–settlement curves of
distances between the test and reaction piles are provided in phase II for helical piles LS12 and LD12.
Fig. 2. The test load was transferred to the reaction piles using Figure 7 shows that the maximum applied load ranged from
high-strength all-threaded steel bars 38 mm in diameter and 994 to 1745 kN for the 6.0 m helical piles tested in phase I and from
1.22 m in length. The main reaction beam was placed on and 2057 to 2552 kN for the 9.0 m helical piles tested in phase II. The
connected to the top of the two secondary beams and centered maximum settlement for helical piles varied from about 43 to
over the test helical pile. For the driven pile, the main reaction 61 mm. The maximum applied load was 811 kN for the driven pile

Published by NRC Research Press


230 Can. Geotech. J. Vol. 52, 2015

Fig. 5. Schematic diagram of test piles: (a) SS11; (b) SD11; (c) SD21; (d) SD31; (e) LS12; (f) LD12.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by University of Western Ontario on 02/17/15
For personal use only.

Published by NRC Research Press


Elkasabgy and El Naggar 231

Fig. 6. Axial compression load test setups: (a) helical pile, schematic view; (b) driven pile, schematic view; (c) helical pile, four reaction piles;
(d) driven pile, two reaction piles. LDT, linear displacement transducer.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by University of Western Ontario on 02/17/15
For personal use only.

tested in phase I, while the maximum settlement reached was slightly increased loads. However, pile load tests do not always
about 41 mm. It can be noted that the driven pile showed initial exhibit a well-defined peak load due to practical limitations on
stiffer response than piles SS11, SD11, and SD21 up to the ultimate loading equipment and test setup. The load–settlement curves
capacity. obtained from the axial compression tests can normally be sim-
The installation of helical and driven piles is expected to cause plified into three distinct regions (Fig. 8): initial linear–elastic
large deformations within the annular zone surrounding the pile. region with steep slope; transition nonlinear region where the
This would result in some reduction in the soil shear strength movements are largely disproportional to the load increments;
within this annular zone. The actual degree of strength reduction and final linear region that shows small stiffness behaviour.
and extent of annular zone are not well defined. It is expected,
Point (L1) in Fig. 8 corresponds to the end of the initial linear
however, that the soil shear strength immediately after installa-
region, and point (L2) represents the onset of the final linear re-
tion would be close to the remoulded strength. As the time passes
gion. Most of the ultimate load interpretation methods define the
after pile installation, it is expected that the soil regains some of
its shear strength. ultimate capacity of piles to fall within the transition nonlinear
region or rarely within the final linear region. Hirany and
Pile ultimate capacity Kulhawy (2002) stated that creep displacements become signif-
Conceptually, the ultimate (or failure) load of a single pile is icant at loads that lie after the transition nonlinear region,
reached when rapid displacements occur under sustained or which can cause difficulty in maintaining a constant load on

Published by NRC Research Press


232 Can. Geotech. J. Vol. 52, 2015

Fig. 7. Axial compression load at pile head versus settlement: (a) helical piles (phase I); (b) driven pile (phase I); (c) helical piles (phase II).
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by University of Western Ontario on 02/17/15
For personal use only.

pile. Consequently, the displacements measured during this Several definitions for the ultimate load of piles have been pro-
loading zone would lead to ultimate load that is subjected to posed to interpret the pile load carrying capacity from its load test
error due to the fluctuating applied load. Thus, it is desirable to data based on mathematical rules for different types of piles
obtain the ultimate load of the pile from the upper limit of the (Fellenius 1980). Terzaghi (1942) defined the ultimate load as the
transition nonlinear zone and before the onset of the final load at which the pile settlement exceeds 10% of its diameter.
linear region. This criterion is expected to overestimate the capacity of large-

Published by NRC Research Press


Elkasabgy and El Naggar 233

Fig. 8. Typical load–settlement curve for drilled foundation. D (i.e., eq. (6)). This method was adopted in the current study for
interpreting the ultimate capacity of helical piles only. This defi-
nition was considered suitable for slender helical piles, as the
helix bearing capacity represents a considerable component of
the pile capacity.

QL
(6) Sp ⫽ ⫹ 0.08D
EpA

Inspecting the load–settlement curves in Fig. 7, it is noted that


the load–settlement curves of the test piles showed the three
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by University of Western Ontario on 02/17/15

distinctive regions (initial linear, transition nonlinear, and final


linear) within the induced settlements. Generally, the near-linear
failure region was achieved at a net settlement of >3.5% of the pile
helix diameter for helical piles (i.e., 0.035D) and 3.5% of the pile
shaft diameter for the driven pile (i.e., 0.035d). Thus, the ultimate
load is defined as the load applied at the pile head, resulting in
total settlement at the pile head equal to 0.035D plus the elastic
deflection of the pile, i.e.,
diameter piles and some pile types. In this study, the ultimate load
capacities of the test piles were established using four commonly QL
(7) Sp ⫽ ⫹ 0.035D
used interpretation methods: Davisson’s method (Davisson 1972); EpA
tangent intersection method (Mansur and Kaufman 1956);
O’Neill and Reese (1999) method; and Livneh and El Naggar (2008)
where, for the driven pile, the helix diameter is replaced with the
method. These interpretation methods aim at locating the pile
ultimate capacity within the transition nonlinear region to reach shaft diameter, d.
an allowable load (design load) that lies in the initial linear region, The application of the developed ultimate capacity criterion is
thus avoiding excessive settlement under operating loads. The illustrated on the load–settlement curves as shown, for example,
For personal use only.

general procedures of these methods are briefly summarized in in Fig. 9 for helical pile SS11 (phase I), driven pile DS (phase I), and
the following text. It should be noted that for any pile that shows helical pile LS12 (phase 2). The 6.0 m helical piles showed a notice-
plunging failure behaviour, the ultimate load is taken as equal to able nonlinear transition zone followed by a linear zone with
the maximum test load. moderate slope (i.e., no sign of plunging failure), reflecting the
effect of the dense to very dense silty sand layer and the clay till
Davisson method (offset limit method) soils where the pile tips rested in. The 9.0 m helical pile LS12,
The Davisson method is primarily intended for driven piles however, displayed almost plunging failure because of the softer
tested according to the quick maintained loading methods. The soil at its tip. In contrast, pile LD12 displayed the initial linear and
failure load is defined as the load corresponding to the pile head the transition nonlinear regions but the final near-failure linear re-
movement that exceeds the pile elastic compression by a value of gion was not reached due to the limitation of the load test setup
4 mm plus a factor equal to the pile diameter divided by 120, i.e., capacity. However, it experienced a higher curvilinear tendency,
which reflected the increased bearing resistance from the helix. The
QL
(5) Sp ⫽ ⫹ (4.0 ⫹ 8d)10⫺3 driven pile, in general, exhibited a more defined ultimate load.
EpA The ultimate load of the test piles were interpreted using the afore-
mentioned methods and the proposed criterion in this study (eq. (7)),
where Sp is the pile head settlement at ultimate load; Q is the and the results are presented in Table 3. It can be concluded that a
applied load; L is the pile length; Ep and A are the pile Young’s considerable scatter exists for the load capacities evaluated using
modulus and cross-sectional area, respectively; and d is the tip different interpretation criteria. The table summarizes the load ca-
diameter, equal to the helix diameter for helical piles and the pacity ratio for the piles, which is defined as the ratio of the differ-
shaft diameter for driven piles. ence between the load capacity obtained from each of the different
ultimate load criteria and the present study’s criterion to that estab-
Tangent intersection method lished from the present study criterion.
The ultimate load capacity is taken as the load associated with The current study failure criterion estimated ultimate capaci-
the intersection of two tangents to the load–settlement curve, one ties varying between 721 and 1331 kN for the 6.0 m piles (phase I),
represents the elastic range and the other represents the progres- and 1952 and 2477 kN for piles LS12 and LD12 (phase II). The cor-
sive settlement. responding pile settlements were between 13.2 and 25.6 mm for
the 6.0 m piles, and between 30 and 33 mm for piles LS12 and
O’Neill and Reese (1999) method
LD12. The ultimate capacities estimated by Davisson criterion var-
The method defines the ultimate axial compression capacity of
ied between 629 and 884 kN for the 6.0 m piles and between 1847
drilled shafts as the load that produces a pile head settlement of
and 1922 kN for piles LS12 and LD12, with corresponding settle-
5% of the pile tip diameter. The tip diameter is taken equal to the
helix diameter for helical piles and the shaft diameter for driven ments that varied between about 10 and 18 mm. However, the
piles. capacities estimated using the tangent intersection criterion were
between 645 and 900 kN for the 6.0 m piles and between 1808 and
Livneh and El Naggar (2008) method 1757 kN for piles LS12 and LD12, with corresponding settlements
The interpretation method is intended for slender helical piles that varied between about 8 and 18 mm. It can be noted that
with small diameter helices. The ultimate load is defined as the Davisson and tangent intersection criteria significantly underes-
load corresponding to pile head movement that exceeds the elas- timated the ultimate capacities of helical piles by about 18%–34%
tic compression of the pile by 8% of the pile largest helix diameter, and 14%–43%, respectively. However, the difference decreased for

Published by NRC Research Press


234 Can. Geotech. J. Vol. 52, 2015

Fig. 9. Developed ultimate capacity criterion: (a) helical pile SS11 the cases of the 9.0 m helical pile, LS12, and the driven pile DS;
(phase I); (b) driven pile DS (phase I); (c) helical pile LS12 (phase II). possibly because these piles have a more defined ultimate load.
The ultimate capacities estimated using O’Neill and Reese cri-
terion varied between 751 and 1478 kN for the 6.0 m piles and
about 1952 and 2477 kN for piles LS12 and LD12, with correspond-
ing settlements of about 31 mm for helical piles and 16 mm for the
driven pile. It can be seen that the O’Neill and Reese criterion
overestimated the ultimate capacities by about 6%–12% for the
6.0 m helical piles, with lesser difference for the driven pile, DS1.
The ultimate capacities estimated by the Livneh and El Naggar
criterion, proposed for slender helical piles, were about 994–
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by University of Western Ontario on 02/17/15

1623 kN for the 6.0 m piles, and about 2057 kN for pile LS12, with
corresponding settlements between about 52 and 59 mm. The
criterion highly overestimated the ultimate capacities of the
tested helical piles by about 18%–44%, but the difference was
5% for pile LS12. Furthermore, it exceeded the maximum load
reached at the end of the load test of piles SD11 and LD12. Conse-
quently, it can be concluded that the Davisson, tangent intersec-
tion, Livneh and El Naggar criteria are not suitable to estimate the
ultimate compression capacity of large-diameter helical piles, es-
pecially those with helices resting in very stiff to hard soils.
As presented in Table 3, the ultimate capacities for the 6.0 m
helical piles varied between 845 and 1331 kN, while the capacity of
the 6.0 m driven pile was 721 kN. The ultimate capacity of the
single-helix pile, SS11, was about 18% higher than that of the
driven pile, DS. However, the ultimate capacity of the double-
helix piles (SD11, SD21, and SD31) was about 45%–85% higher than
that of the driven pile. The ultimate capacity of the helical piles
occurred at pile head displacement values that were ⬃85% higher
For personal use only.

than the displacement at the ultimate capacity of the driven pile.


This can be attributed to the beneficial contribution of the large
diameter helices on the mobilized end bearing capacity. Gener-
ally, it demonstrates the advantage of helical piles over driven
piles in ground conditions similar to what was encountered in the
test site. However, the range of ratios of axial compression capac-
ity of the helical piles to the driven pile reported herein (i.e.,
1.18–1.85) is smaller than the range reported by Sakr (2011) as
3.0–5.0.
Figure 10 depicts the applied load normalized by the ultimate
capacity (Q /Q ult) and the settlement, St, normalized by the helix
diameter, D, for helical piles and by the shaft diameter, d, for the
driven pile. It can be noted from Fig. 10a and capacity values of
the 6.0 m helical piles that the silty sand layer that existed at the
elevation of the pile tip increased the pile capacity and eliminated
the plunging failure trend.
The site investigation program showed that the silty sand layer
varies in thickness along the test site and in some locations may
not exist in the whole interhelix zone for all the 6.0 m helical
piles. For some piles, the interhelix zone was located within the
silty clay and silty sand layers. In addition, the strength and stiff-
ness of the clay till layer varied along the test site and with depth.
A higher degree of soil disturbance is expected to occur due to
installation under the upper helices compared to the bearing soil
under the lower helices, which can significantly compromise
their end bearing resistance. These factors affected the general
trend of the ultimate capacity of the 6.0 m helical piles.
The pile design capacity is usually evaluated as one-half of its
ultimate capacity as established from the load test. The load–
settlement curves show that the pile settlement at the design
capacity is quite small. The measured settlements at the design
capacities for all piles tested in this study were <1% of the helix
diameter for the helical piles and <1% of the pile shaft diameter
for the driven pile. The settlement at the pile design capacity was
slightly less for the driven pile compared to the helical piles.
However, the settlement of the driven pile at the same load value
is greater than that of the helical pile. This demonstrates that the
performance of helical piles is superior to that of the driven pile at
the same load level.

Published by NRC Research Press


Elkasabgy and El Naggar 235

Table 3. Ultimate pile capacities, settlements, and load capacity ratio established from different criteria.
Davisson Tangent intersection O’Neill and Reese Livneh and El Naggar Present study criterion
Load (kN) Load (kN) Load (kN) Load (kN)
Pile No. (Ratio %)* Settl. (mm) (Ratio %)* Settl. (mm) (Ratio %)* Settl. (mm) (Ratio %)* Settl. (mm) Load (kN) Settl. (mm)
Phase I (piles tested 2 weeks after installation)
SS11 629 (−26) 11 731 (−13) 18.2 894 (6) 30.5 994 (18) 51.8 845 24.4
SD11 884 (−34) 11.5 840 (−37) 10.4 1478 (11) 30.5 —** —** 1331 25.0
SD21 778 (−31) 11.8 645 (−43) 7.5 1259 (12) 30.5 1623 (44) 54.0 1126 24.8
SD31 860 (−18) 11.9 900 (−14) 13.7 1116 (6) 30.5 1281 (22) 53.0 1049 24.5
DS 691 (−4) 10 677 (−6) 9.3 751 (4) 16.0 —** —** 721 13.5
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by University of Western Ontario on 02/17/15

Phase II (piles tested 9 months after installation)


LS12 1847 (−5) 17.7 1808 (−7) 16.8 1952 (0) 30.5 2057 (5) 58.6 1952 30.3
LD12 1922 (−23) 17.9 1757 (−29) 15.2 2477 (0) 30.5 —** —** 2477 32.7
Note: Settl., settlement.
*Load capacity ratio (values in parentheses) defined as the ratio (%) of the difference between load capacity obtained from each of the different ultimate load criteria
and the present study's criterion to that established from the present study criterion.
**Ultimate capacity could not be obtained by the Livneh and El Naggar criterion.

Distribution of axial load along helical piles and around the helices. Thus, the bearing resistance of the upper
The capacity of helical piles is developed from three load- helix may depend to some extent on the shear strength of the soil
transfer mechanisms, namely resistance along the shaft above the located outside the disturbed soil in the interhelix zone.
helices, resistance along the interhelix zone, and resistance at the The contribution of each load-transfer mechanism is calculated
bottom helix. To establish the contributions of these load-transfer from the load-transfer curves, and the results are illustrated in
mechanisms, the distribution of the axial compression load along Fig. 13. These figures show the mobilized skin friction and end
helical piles during the load test was evaluated from the analysis bearing resistances at the ultimate capacity estimated by the
of the strain gauge readings at different elevations along the pile. criterion proposed in the present study. The dotted line, which
Since the driven pile, DS, had no strain gauges, the load distribu- indicates the ultimate capacity, was drawn at a total pile head
For personal use only.

tion for this pile was not included in this section. settlement equal to the elastic pile shortening plus a net settle-
The load distribution curves at various load levels were estab- ment of 0.035D. The figure demonstrates that the initial shaft
lished and are shown in Figs. 11 and 12. For single-helix piles, the response was very stiff, and its capacity was mobilized at a settle-
ultimate capacity is made of shaft friction and end bearing resis- ment of about 1.0% of the shaft diameter and 0.55% of the helix
tance from the bottom helix – pile tip. For double-helix piles, the diameter. In addition, the end bearing underneath the helices –
ultimate capacity is composed of four components: shaft friction, pile tip was very small initially, and was mobilized fully at a much
bearing resistance on the upper helix, shaft friction along the larger settlement. It is important to note that shaft resistance was
interhelix zone, and bearing resistance from the bottom helix – mobilized fully, and the ultimate load defined by the proposed
pile tip. It can be seen from the load distribution curves that the criterion generally is defined at the end of the curvilinear region
double-helix piles experienced the individual bearing behaviour of the end bearing resistance curve of the bottom helix – pile tip.
as the upper helix mobilized a considerable value of bearing re- This confirms the adequacy of the proposed criterion to estimate
sistance. the ultimate capacity of large-capacity helical piles.
By examining Figs. 11b and 12b, it is interesting to note that piles The percentage contributions of the different components at
SD11 and LD12 with interhelix spacing, S/D = 1.5, showed mainly the ultimate capacity, Q ult, and at the design load (0.5Q ult) are
individual plate bearing failure mechanism instead of the cylin- summarized in Table 4. Careful examination of the load distribu-
drical shear failure. Previous studies by Narasimha Rao et al. tion curves and the analysis of load-transfer mechanism results
(1989), Narasimha Rao et al. (1991), Narasimha Rao et al. (1993), and presented in Table 4 yield the following observations:
Zhang (1999) examined the effect of S/D on the failure mechanism
of helical piles of small shaft diameter under compression and • The percentage of shaft capacity and interhelix capacity de-
uplift loads. These studies showed that the cylindrical shear fail- creased as the applied load increased because most of the fric-
ure was predicted for helical piles in clay soil with S/D ≤ 1.5–2.0, tional capacity was mobilized within the first 3–4 mm of pile
while for S/D > 2.0 the individual plate bearing failure mechanism settlement. In contrast, the percentage of helices bearing ca-
would dominate. pacity increased as the applied load increased because higher
The observed behaviour in the current study reflects the distur- settlement is required to mobilize the end bearing component.
bance of structured glacial clayey and silty deposits within the It is also noted that the end bearing resistance continued to
interhelix zone due to pile installation. The behaviour of piles increase with the increase of the applied load even after the
SD11 and LD12 suggests that the upper helix transferred load to estimated ultimate capacity, especially for piles that showed a
the disturbed soil underneath it, resulting in substantial deforma- significant curvilinear tendency in the load–settlement curves.
tions before there was sufficient displacement of the interhelix • The percentage of shaft resistance decreased from 11% to 2%, for
soil relative to the intact soil outside the helices zone to mobilize the 6.0 m helical piles as the embedment ratio, H/D, decreased
cylindrical shear resistance. This behaviour happened to a lesser from 8.9 to 4.5. However, the adhesion in the interhelix section
degree in the case of pile LD12, as the pile was tested 9 months increased from 1.3% to 3.1% as the interhelix spacing ratio in-
after installation, allowing some regain of soil strength. creased from 1.5 to 4.5. A slightly higher shaft resistance (12.2%)
For the double-helix piles SD21 and SD31 with widely spaced was mobilized for the 9.0 m helical piles.
helices (S/D = 3.0 and 4.5, respectively), a plastic wedge of soil is • All double-helix 6.0 m piles experienced individual plate bear-
expected to develop underneath each helix. To mobilize pile fail- ing failure mechanism with the upper helix contribution of
ure, a soil mass located outside the plastic wedge, and outside the 6%–10% of the ultimate capacity. This is attributed to the distur-
disturbed zone, would be displaced and deformed. This soil mass bance of the structured interhelix glacial clayey and silty depos-
would not be as large as in the case of undisturbed soil beneath its due to pile installation.

Published by NRC Research Press


236 Can. Geotech. J. Vol. 52, 2015

Fig. 10. Normalized load – settlement curves: (a) helical piles • The double-helix 9.0 m pile LD12 experienced individual plate
(phase I); (b) driven pile (phase I); (c) helical piles (phase II). bearing failure with the upper helix contribution of 22.5% of
the ultimate capacity.
• The bearing resistance contribution of the bottom helix – pile tip
was about 85%–90% of the ultimate capacity of the 6.0 m piles
(phase I) and about 67% and 87% for the 9.0 m piles (phase II).

Bearing pressure of helices compared with CPT resistance


The soil resistance below the helices – pile tip and the soil resis-
tance to the advancement of cone penetrometer are functions of the
soil shear strength. In addition, the loading mechanism is similar in
both cases. Thus, the ultimate bearing pressure below the helices –
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by University of Western Ontario on 02/17/15

pile tip may be correlated to the cone tip resistance.


Figure 14 plots the profiles of corrected cone tip resistances, qt,
measured on site at locations SCPT-1, SCPT-2, and SCPT-3 along
with the bearing pressures below the helices and pile tip. It is
noted that the bearing pressures below the upper helix of the
double-helix 6.0 m piles tested 2 weeks after installation (phase I)
were significantly less than the range of cone resistance measure-
ments. This is likely due to the soil disturbance caused by pile
installation. Furthermore, the bearing pressures below the bot-
tom helix – pile tip of the same 6.0 m piles were significantly less
than the CPT resistance measured in the silty sand layer, probably
due to the variability of the thickness of silty sand layer. By com-
paring profiles of SCPT-1, SCPT-2, and SCPT-3, it can be seen that
the silty sand layer was shallower towards the west side of the test
site (see Figs. 2 and 4).
The bearing pressure below the bottom helix – pile tip of the
9.0 m piles tested 9 months after installation (phase II) was higher
For personal use only.

than the CPT resistance, but the bearing pressure below the upper
helix of pile LD12 was approximately similar to some extent to the
lower value of the cone resistance profiles. This is ascribed to the
thixotropic behaviour of the soil, as it regained most of its shear
strength with time after pile installation and the existence of
interbedded layers of silty sand within the clay till layer. Accord-
ingly, the cone tip resistance along with an adequate soil investi-
gation program can provide a reasonable estimate for the bearing
resistance of the bottom helix – pile tip.

Inferred soil resistance mobilized by helical piles


The average values of unit shaft adhesion along the upper 4.0 m
of pile shaft were calculated for all test helical piles at the ultimate
loads, and the results are presented in Table 5. It is observed that
the mobilized shaft adhesion value in the stiff clay to silty clay
layer (layer 2) for pile SS11 was 37% higher than the corresponding
values calculated for the other 6.0 m piles. This may be ascribed to
the possible higher disturbance induced by the multi-helix piles
and the potential effect of layered soil medium on increasing
disturbance as discussed earlier. In addition, a significant differ-
ence of about 70%–135% exists between the mobilized unit shaft
adhesion values along the stiff clay to silty clay layer (layer 2) for
piles SD11, SD21, and SD31. This is attributed to the minimal ad-
hesion developed over the shaft length immediately above the
upper helix. In the current study, the shaft adhesion was ne-
glected along a length equal to one helix diameter, D, above the
upper helix, and the corrected unit shaft adhesion values were
calculated for piles SD11, SD21, and SD31 and provided in Table 5.
The revised unit shaft adhesion values along the stiff clay to silty
clay layer (layer 2) are in agreement at different depths along the
upper 4 m of pile shaft. It should be noted that no correction was
required for the unit shaft adhesion values for piles SS11, LS12, and
LD12 along the upper 4 m, as the helix was deep enough below
strain gauges of level SG4 (see Table 5). Similar results were re-
ported by Zhang (1999). The researchers suggested that the shaft
adhesion could not be mobilized along a length of one helix di-
ameter, D, above the upper helix because of the shadowing effect.
It was proposed to use an effective shaft length for calculating the
shaft adhesion, given as Heff = H − D.

Published by NRC Research Press


Elkasabgy and El Naggar 237

Fig. 11. Load distribution curves for piles tested in phase I: (a) SS11; (b) SD11; (c) SD21; (d) SD31.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by University of Western Ontario on 02/17/15
For personal use only.

The unit shaft adhesion values along the silt crust (layer 1) and The formulations and correlations provided by Bishop (1966),
in the stiff clay to silty clay (layer 2) for piles LS12 and LD12 tested Kulhawy and Mayne (1990), Mitchell (1993), Stark and Eid (1994),
9 months after installation are significantly higher than the cor- and Terzaghi et al. (1996) showed that the angle of internal fric-
responding values for the 6.0 m helical piles tested 2 weeks after tion, ␾ , of the glacial clay till was found to vary between approx-
installation. The unit shaft adhesion values increased in average imately 28° and 35°. Thus, it was deemed reasonable to use an
by about 16%–50% for the silt crust and between about 54% and average value of ␾  = 31° for the glacial clay till layer.
130% for the stiff clay to silty clay layer. Cemented and heavily overconsolidated clays mobilize effec-
The undrained and drained soil cohesion, Cu and C=, respec- tive cohesion, C=, under loading (Kulhawy and Mayne 1990). Mesri
tively, and the effective angle of internal friction, ␾ , mobilized by and Abdel-Ghaffar (1993) related the short-term value of C= to the
the different sections of the helical piles, were back-calculated preconsolidation pressure, ␴p , and the current stress state, i.e.,
using the design equations given in the section “Individual plate C= = 0.024␴p . This relationship gives an effective cohesion for the
bearing method” and the obtained load-transfer curves at ulti- glacial clay till layer that varies between 18 and 26 kPa. Thus, a C=
mate loads, employing a trial-and-error technique. value of 20 kPa is considered reasonable in the analysis.

Published by NRC Research Press


238 Can. Geotech. J. Vol. 52, 2015

Fig. 12. Load distribution curves for piles tested in phase II: (a) LS12; (b) LD12.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by University of Western Ontario on 02/17/15
For personal use only.

The average undrained cohesion mobilized along the pile shaft in Table 6 provides the back-calculated mobilized shear strength
the different silty clay and clayey silt layers was calculated using the parameters for the 6.0 m helical piles tested 2 weeks after instal-
shaft diameter, d = 0.324 m, and an adhesion factor, ␣, that varied lation. The results show that the soil disturbance had a significant
between 0.4 and 1.0 as per the recommended values by the Canadian effect on the mobilized undrained and drained shear strength
Foundation Engineering Manual (CFEM 2006), such that parameters along the shaft and upper helix sections. The back-
calculated ␾  for the silt crust (layer 1) was about 75%–88% of the
(8) Cu ⫽ Q shaft / ␣␲dHshaft intact value obtained from the site investigation program (␾  =
31o). The mobilized undrained cohesion, Cu, of the stiff clay – silty
clay and clayey silt (layer 2) was about 20%–45% of the shear
where Hshaft is the length of the shaft section corresponding to a
strength of the intact layer. In contrast, the reduction in the un-
certain soil layer, and Q shaft is the ultimate pile shaft capacity. The
drained cohesion of the very stiff silty clay – clayey silt – clay (layer
mobilized drained shear strength parameters of the silty crust
3) was about 60%.
and clay till were obtained as follows:
Table 7 provides the results for the 9.0 m helical piles, LS12 and
LD12, tested 9 months after installation. For the silt crust (layer 1),
␲ 2
(9) Q shaft ⫽ ␣C ␲dHshaft ⫹ dH ␥ Ks tan␦ the mobilized ␾  was about 17%–42% higher than the values cal-
2 shaft
culated for piles tested in phase I. The mobilized Cu values in-
creased by about 45%–250% (in layer 2) and by about 53%–67% (in
where Ks and ␦ are the coefficient of lateral earth pressure and layer 3) relative to the values mobilized for piles tested in phase I.
pile–soil friction angle, respectively, obtained from Kulhawy The mobilized drained shear strength parameters, C= and ␾ , of
(1984) and the CFEM (2006). the clay till layer were in average of about 70% and 74% of the
The average undrained cohesion mobilized by the upper helix interpreted intact values, respectively. However, it can be con-
for all 6.0 m double-helix piles was back-calculated considering cluded that the soil disturbance occurred in the soil during instal-
the individual plate bearing method, and using the bearing capac- lation persisted long after installation, which was illustrated in
ity factor, Nc = 9.0, as follows: comparing the mobilized shear strength parameters 9 months
after pile installation with the intact ones due to the structured
(10) Q upper-helix ⫽ NcCuAh nature of the soils encountered in the test site.
It should be noted that the induced effective stresses in the site
where Q upper-helix is the ultimate bearing capacity of the upper soils during the penetration of the pile shaft and helices exceeded
helix, and Ah is its projected area. The average drained shear their yield stresses, which resulted in significant vertical and ra-
strength parameters mobilized by the upper helix for the 9.0 m dial strains that caused the interparticle bonds to break and ex-
double-helix pile, LD12, were back-calculated using Nc = 9 and Nq, cess pore-water pressure to generate due to the collapse of the soil
considering the Meyerhof (1976) approach but using a factor of structure. Therefore, the soil adjacent to the pile shaft and helices
safety of 2.0, as follows: experienced a significant drop in its shear strength that was man-
ifested in the lesser mobilized shaft unit friction resistances and
(11) Q upper-helix ⫽ Ah共C Nc ⫹ ␴voNq兲 strength parameters after installation. As such, Roy and Lemieux
(1986), through in situ tests, indicated that the soil memory of clay
soil adjacent to driven piles would be erased due to soil remould-
where ␴vo is the effective overburden pressure. ing.

Published by NRC Research Press


Elkasabgy and El Naggar 239

Fig. 13. Load transferred to soil for piles: (a) SS11; (b) SD11; (c) SD21; (d) SD31; (e) LS12; (f) LD12.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by University of Western Ontario on 02/17/15
For personal use only.

Published by NRC Research Press


240 Can. Geotech. J. Vol. 52, 2015

Table 4. Percentage of load transferred by different sections of helical the pile shaft and the surrounding soil due to undetected minor
piles. imperfections during the welding of the pile shaft segments.
Pile After pile installation, the effective stress of soil adjacent to the
pile increases due to the dissipation of the excess pore pressure
Load Pile section SS11 SD11 SD21 SD31 LS12 LD12 accompanied by a reduction in the void ratio. The reconsolidation
0.5Q ult Shaft 20.3 7.2 6.4 4.9 23.6 17.0 of soil causes the yield stress to increase in step with the increase
0.5Q ult Upper helix — 11.1 4.7 10.9 — 19.4 in the effective stress, leading to an increase in the strength pa-
0.5Q ult Interhelix — 2.6 5.0 5.2 — 1.6 rameters of soil layers. Further increase in the strength and stiff-
0.5Q ult End bearing* 79.7 79.1 83.9 79.0 76.4 62.0 ness of the reconsolidated soils could be achieved by the processes
Q ult Shaft 11.0 4.1 3.3 2.0 12.7 9.4 of aging and thixotropic hardening. The aging process employs a
Q ult Upper helix — 9.7 10.0 5.9 — 22.5 decrease in volume of soil and an increase in the vertical yield
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by University of Western Ontario on 02/17/15

Q ult Interhelix — 1.3 1.6 3.1 — 1.2 stress under a constant effective vertical stress. An additional in-
Q ult End bearing* 89.0 84.9 85.1 89.0 87.3 66.9 crease in the vertical yield stress and hence the soil strength may
*End bearing resistance from bottom helix and pile tip. occur due to the thixotropic hardening phenomenon in which the
thixotropic bonding between the soil particles takes place, when
Fig. 14. Bearing pressures at failure generated by helices and pile the rate of change in void ratio is slow enough. The thixotropic
tip compared with cone tip resistances (qt). cohesive soil would have the tendency to regain the interparticle
force balance and reorganize the adsorbed water–cation structure
to a lower energy state. It should be noted that Mitchell (1993)
indicated that thixotropic hardening may account for low to me-
dium sensitivity of the cohesive soils.
The mobilized undrained cohesion values underneath the up-
per helix embedded in layer 2 were about 70%–74% of the undis-
turbed values and about 50% for the case of the upper helix
embedded in layer 3. These high percentages are probably due to
the propagation of the failure surface outside the disturbed zone
into the undisturbed soil beyond the helix diameter, or propaga-
tion of failure surface into the underlying very stiff silty clay –
clayey silt (layer 3). For the upper helix section in pile LD12, the
For personal use only.

mobilized C= and ␾  values of clay till were about 90% and 93.5% of
the intact parameters, respectively. These high percentages are
attributed to the propagation of the failure surface outside the
disturbed zone.
The ultimate end bearing capacities mobilized by the bottom
helix – pile tip were calculated using the intact shear strength
parameters and compared with the measured values, as it was
believed that pile installation process would introduce minimal
disturbance below the bottom helix – pile tip. Due to the variabil-
ity in the thickness of the silty sand layer located near the 6.0 m
pile tip, the shear strength parameters of the clay till layer and the
silty sand layer were used to calculate the end bearing resistances
of the bottom helix – pile tip for all 6.0 m piles (SS11, SD11, SD21,
and SD31). The ultimate bearing capacity of the bottom helix – pile
tip was calculated as follows:

(12) Q Bottom ⫽ (Ah ⫹ At)共C Nc ⫹ ␴voNq兲

where At is the area of the pile tip, and the bearing capacity factors
Nc and Nq were obtained from Meyerhof (1976) using a safety factor
of 2.0 (assuming behaviour at helical pile tip is similar to bored
The estimated overconsolidation ratios of the test site soils in- piles). The average calculated ultimate end bearing capacity for
dicate moderately and highly overconsolidated soil layers. It is the 6.0 m piles considering clay till was 872 kN (i.e., equivalent to
believed that such soils were subjected to aging and hardening the end bearing pressure of about 3000 kPa). The measured values
during the geological process of the natural development of the varied between 752 and 1131 kN. Thus, the percentage difference
soils’ microstructure. However, they experienced a decrease in between the predicted capacity and the measured capacity varied
their strength and stiffness parameters when those geologic ef- between 15.9% and −22.9%. Therefore, the ultimate capacity of the
fects were removed due to the destruction of the soil microstruc- bottom helix – pile tip can be estimated using the intact shear
ture. Perko (2009) observed decreased pile capacities for helical strength parameters. Furthermore, Fig. 14 illustrates that the
piles installed and tested in highly overconsolidated soils. Fur- measured end bearing pressures below the bottom helix – pile tip
thermore, Tappenden (2007), after conducting several load tests of the 6.0 m piles are in close agreement with the cone tip
on helical piles installed in typical Alberta soils, concluded that resistance. In contrast, the average calculated ultimate bearing
disturbance to the native soils during the pile installation could resistance for the 6.0 m piles using the drained shear strength
severely reduce the expected capacity for piles installed in highly parameter of the silty sand layer (␾  = 42°) was 2565 kN (i.e.,
overconsolidated or structured soils. The reduction in the soil equivalent to the end bearing pressure of 8780 kPa), considering
strength parameters as inferred from the measured shaft resis- the Meyerhof (1976) limiting point resistance of q l = 0.5Nq tan(␾ ),
tances can also be ascribed to testing the phase I piles during the where q l is in t/ft2 (1 ft = 0.3048 m). This significantly overesti-
wet season and probable formation of very small gaps between mated the capacity by 127%–241%. Moreover, Fig. 14 shows that the

Published by NRC Research Press


Elkasabgy and El Naggar 241

Table 5. Unit shaft adhesion for helical piles along upper 4.0 m of pile shaft.
Strain gauge depth (m) Uncorrected unit Corrected unit
Pile (stain gauge level) Soil layer shaft adhesion (kPa/m) shaft adhesion (kPa/m)
SS11 0–1.35 (SG1–SG2) Layer 1, silty to sandy silt 6.2 6.2
1.35–2.85 (SG2–SG3) Layer 2, stiff clay to silty clay 24.0 24.0
2.85–4.35 (SG3–SG4) Layer 2, stiff clay to silty clay 24.2 24.2
SD11 0–1.60 (SG1–SG2) Layer 1, silty to sandy silt 7.7 7.7
1.60–3.10 (SG2–SG3) Layer 2, stiff clay to silty clay 18.0 18.0
3.10–4.50 (SG3–SG4) Layer 2, stiff clay to silty clay 10.6 17.4
SD21 0–1.30 (SG1–SG2) Layer 1, silty to sandy silt 6.6 6.6
1.30–2.45 (SG2–SG3) Layer 2, stiff clay to silty clay 17.1 17.1
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by University of Western Ontario on 02/17/15

2.45–3.0 (SG3–SG4) Layer 2, stiff clay to silty clay 7.3 15.2


SD31 0–1.60 (SG1–SG2) Layer 1, silty to sandy silt 5.9 5.9
1.60–2.73 (SG2–SG3) Layer 2, stiff clay to silty clay 9.5 17.3
LS12 0–1.5 (SG1–SG2) Layer 1, silty to sandy silt 8.9 8.9
1.5–3.0 (SG2–SG3) Layer 2, stiff clay to silty clay 35.5 35.5
3.0–4.35 (SG3–SG4) Layer 2, stiff clay to silty clay 36.1 36.1
LD12 0–1.5 (SG1–SG2) Layer 1, silty to sandy silt 9.4 9.4
1.5–3.0 (SG2–SG3) Layer 2, stiff clay to silty clay 36.7 36.7
3.0–4.61 (SG3–SG4) Layer 2, stiff clay to silty clay 37.2 37.2
Note: Shaded values indicate revised unit shaft adhesion values.

Table 6. Back-calculated shear strength parameters mobilized by piles in phase I.


Mobilized shear strength Ratio of mobilized to intact
Shaft Upper helix Shaft Upper helix
Layer Cu (kPa) ␾  (o) Cu (kPa) Cu (%) 
␾ (%) Cu (%)
Pile SS11
Layer 1, silt crust — 25.0 — — 80.6 —
For personal use only.

Layer 2, stiff clay to silty clay and clayey silt 38.5 — — 45.3 — —
Layer 3, very stiff silty clay, clayey silt, and clay 54.0 — — 39.4 — —

Pile SD11
Layer 1, silt crust — 27.0 — — 87.1 —
Layer 2, stiff clay to silty clay and clayey silt 17.6 — — 20.7 — —
Layer 3, very stiff silty clay, clayey silt, and clay — — 68.1 — — 50.0

Pile SD21
Layer 1, silt crust — 27.5 — — 88.7 —
Layer 2, stiff clay to silty clay and clayey silt 16.5 — 59.5 19.4 — 70.0

Pile SD31
Layer 1, silt crust — 23.5 — — 75.8 —
Layer 2, stiff clay to silty clay and clayey silt 17.3 — 33.0 20.4 — 38.8

Table 7. Back-calculated shear strength parameters mobilized by piles in phase II.


Mobilized shear strength Ratio of mobilized to intact
Shaft Upper helix Shaft Upper helix
Layer Cu (kPa) C= (kPa) ␾  (o) C= (kPa) ␾  (o) Cu (%) C= (%) 
␾ (%) C= (%) ␾  (%)
Pile LS12
Layer 1, silt crust — — 32.0 — — — — 103.0 — —
Layer 2, stiff clay to silty clay and clayey silt 55.0 — — — — 65.0 — — — —
Layer 3, very stiff silty clay, clayey silt, and clay 82.5 — — — — 60.2 — — — —
Layer 5, clay till — 14.0 23.0 — — — 70.0 74.0 — —

Pile LD12
Layer 1, silt crust — — 33.5 — — — — 108.1 — —
Layer 2, stiff clay to silty clay and clayey silt 58.0 — — — — 68.2 — — — —
Layer 3, very stiff silty clay, clayey silt, and clay 87.0 — — — — 63.5 — — — —
Layer 5, clay till — 14.0 23.0 18.0 29.0 — 70.0 74.0 90.0 93.5

measured end bearing pressures were much less than the cone tip layer, as the thickness of the silty sand layer below the piles was
resistance in the silty sand layer. Hence, it can be concluded that small.
the end bearing resistances mobilized below the bottom helix – The measured values of the ultimate bearing capacity of the
pile tip of the 6.0 m piles were developed mainly from the clay till 9.0 m piles LS12 and LD12 were 1704 and 1658 kN, respectively.

Published by NRC Research Press


242 Can. Geotech. J. Vol. 52, 2015

However, the average calculated ultimate bearing capacity using compression loading and tension loading of piles can also be at-
the intact clay till strength parameters was 1090 kN, representing tributed to the fact that the bottom helix rests on relatively un-
a difference of –36.0% and −34.2% for piles LS12 and LD12, respec- disturbed soil in compression loading, but in tension loading the
tively. In contrast, the Meyerhof theory overestimates the bearing upper helix bears on the soil affected by the installation of the
capacity of the bottom helix – pile tip (2565 kN) by about 50%–55%. helices (i.e., disturbed soil). Hence, it is recommended by the au-
The measured ultimate end bearing resistance below the bottom thors to use the torque–capacity empirical relationship as a qual-
helix – pile tip of the 9.0 m helical piles lies between the ultimate ity assurance measure during pile installation.
capacity calculated using the clay till drained strength parameters
and that calculated using the silty sand drained strength param- Summary and conclusions
eters. This can be ascribed to the interbedding of silty sand and In total, seven axial compression tests were conducted, and the
clay till layers. measured pile axial response and strain values were used to
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by University of Western Ontario on 02/17/15

evaluate the piles’ ultimate capacities and the load-transfer mech-


Installation torque to helical pile capacity anisms. The mobilized shear strength parameters were back-
The relationship between helical pile ultimate capacity and calculated for helical piles tested 2 weeks and 9 months after pile
installation torque has been used as a general rule of thumb installation. Based on the experimental results and their analysis,
in industry since the 1960s (Perko 2009). The empirical torque– the following conclusions can be drawn:
capacity relationship was first developed by Hoyt and Clemence
(1989) for small diameter helical piles or anchors resisting uplift 1. An ultimate capacity criterion was proposed to estimate the
loads. The relationship is also reported in the CFEM (2006). The ultimate load of large-capacity helical piles. The ultimate ca-
empirical relationship can be presented as follows: pacity is defined as the load corresponding to the pile head
settlement that equals the pile elastic deformation plus 3.5%
(13) Q ult ⫽ KTT of the helix diameter. The proposed failure criterion is appro-
priate for large-diameter helical piles, as it identifies the
failure load within the nonlinear region of the pile load–
where KT is the empirical capacity-to-torque factor and has units settlement curve.
of m−1 (ft−1), and T is the average installation torque. A number of 2. The Davisson (1972) criterion and tangent intersection
theoretical capacity-to-torque correlations were developed by sev- method significantly underestimated the ultimate capacities
eral researchers, including Perko (2001), Perko (2009), and Tsuha of the test large-capacity helical piles, while the criterion de-
and Aoki (2010). veloped by Livneh and El Naggar (2008) for slender helical
For personal use only.

Based on the results of the full-scale compression load tests piles significantly overestimated their ultimate capacities.
conducted on large-diameter helical piles in the current study 3. The ultimate axial capacity of the 6.0 m double-helix piles
and the measured torque values, the back-calculated capacity-to- were about 24%–57% higher than that of the single-helix pile
torque factors for the tested piles are presented in Table 1. The of similar pile length and diameter.
capacity-to-torque factors, KT, for piles tested in phase I were 4. The ultimate axial compression capacity of the 9.0 m double-
10.2 m−1 (3.1 ft−1) for the single-helix pile SS11 and between 10.0 m−1 helix pile was about 27% higher than that of the single-helix
(3.0 ft−1) and 11.6 m−1 (3.5 ft−1) for the double-helix piles SD11, SD21, pile of similar pile length and diameter.
and SD31. Moreover, the KT values for the single-helix and double- 5. The ultimate capacity of the 6.0 m single-helix pile was
helix piles, LS12 and LD12, tested in phase II were 20.5 m−1 (6.2 ft−1) higher by about 18% than the capacity of the 6.0 m driven pile,
and 26.5 (8.0 ft−1), respectively, which are higher than those ob- while the ultimate capacities of the 6.0 m double-helix piles
tained in phase I of testing by about 101% for the single pile and by were higher by about 45%–85% than the capacity of the 6.0 m
about 128%–165% for the double-helix piles. It should be noted that driven pile, considering the site subsurface conditions. These
the obtained torque values for piles LS12 and LD12 are considered results confirm the advantage of helical piles compared to the
low for the estimated consistency of the till deposits around the conventional driven piles of similar pile embedded length
pile tip, which resulted in the high KT numbers calculated for and shaft diameter.
those piles. The low torque values may be attributed to the layer- 6. The measured settlements at ultimate compression capacity
ing nature of the site soil, which affects the bearing strata under of helical piles were 23.8–25.6 mm for the 6.0 m helical piles
the helices at different depths or a possible source of error in tested 2 weeks after pile installation, and 30–33 mm for the
measuring the torque. Thus, the torque and KT values for piles 9.0 m helical piles tested 9 months after pile installation. This
LS12 and LD12 should be cautiously considered. difference in settlement is due to the nature of soil under-
The CFEM (2006) recommended the use of a KT value of neath the bottom helices. In the case of the 6.0 m helical
10.0 m−1 (3.0 ft−1) for helical piles with shaft diameter approaching piles, the helices were bearing on layers 2 and 3 (stiff silty
200 mm. Tappenden (2007) reported a KT value of 9.19 m−1 for piles clay, clayey silt, and clay), while for the 9.0 m piles, the helices
of shaft diameter between 140 and 400 mm installed in western were bearing on layer 5 (clay till). The till withstood larger
Canadian soils and loaded in compression or tension. For piles of settlement before failure occurred. The settlement at failure
shaft diameter 267 mm, installed in oil sands, and loaded in com- in the case of the 6.0 m driven pile, tested 2 weeks after pile
pression, Sakr (2009) reported a KT of 23.9 m−1. It can be seen that installation, was 13.6 mm. This is because the driven pile has
the calculated KT values for the piles tested in phase I are in good derived most of its capacity through shaft resistance, while
agreement with that recommended by the CFEM (2006) and helical piles have derived the majority of their capacities
slightly higher than that reported by Tappenden (2007). through end bearing (see Table 4). It is well known that end
It should be mentioned that the installation torque is depen- bearing resistance requires large settlement to mobilize.
dent on several factors such as soil strength, existence of large- 7. The measured settlements at the design capacities (i.e., by
size coarse particles (i.e., such as cobbles and boulders), pile shaft employing a factor of safety of 2.0) for all piles tested in this
diameter and shape (i.e., circular or square shaft), helix diameter, study were <1% of the helix diameter for the helical piles
pitch size, number of helices, applied compression load on pile and <1% of the pile shaft diameter for the driven pile.
during installation, etc. In addition, the recommended KT values 8. The tested double-helix piles showed mainly individual plate
reported in the literature were mainly for piles under tension bearing failure mechanism, even for piles with interhelix
loading or did not differentiate between piles under compression spacing of 1.5 times the helix diameter, contrary to the com-
or tension loads. Thus, such differences between KT obtained for mon cylindrical shear failure mechanism expected for such

Published by NRC Research Press


Elkasabgy and El Naggar 243

interhelix spacing ratio. This is attributed to the high degree finite element method. Canadian Geotechnical Journal, 46(6): 627–638. doi:
of soil disturbance and softening of the structured site soils. 10.1139/T09-008.
Livneh, B., and El Naggar, M.H. 2008. Axial testing and numerical modeling of
9. The end bearing capacity of the bottom helix – pile tip could square shaft helical piles under compressive and tensile loading. Canadian
be reasonably predicted using the cone tip resistance profiles Geotechnical Journal, 45(8): 1142–1155. doi:10.1139/T08-044.
measured during the cone penetration testing in the intact Lutenegger, A.J. 2009. Cylindrical shear or plate bearing? – Uplift behavior of
soil and also from using the Meyerhof bearing capacity the- multi-helix screw anchors in clay. Contemporary Issues in Deep Foundations,
ory for deep foundations. ASCE, 2009: 456–463. doi:10.1061/41021(335)57.
10. The mobilized shaft shear resistance of the helical piles de- Mansur, C.L., and Kaufman, J.M. 1956. Pile tests, low-sill structure, Old River,
Louisiana. Journal of Soil Mechanics and Foundation Division, 82(SM5): 1–33.
creased due to soil disturbance associated with pile installa-
Merifield, R.S., Laymin, A.V., and Sloan, S.W. 2006. Three-dimensional lower-
tion. The mobilized strength parameters of stiff to very stiff bound solutions for the stability of plate anchors in sand. Géotechnique,
silty clay and clayey silt soils were about 20%–40% of the 56(2): 123–132. doi:10.1680/geot.2006.56.2.123.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by University of Western Ontario on 02/17/15

intact values. However, the mobilized shaft resistance in- Mesri, G., and Abdel-Ghaffar, M.E.M. 1993. Cohesion intercept in effective stress-
creased significantly with time (during the 9 month period) stability analysis. Journal of Geotechnical Engineering, 119(8): 1229–1249.
to about 65% of the intact values for stiff to very stiff clayey doi:10.1061/(ASCE)0733-9410(1993)119:8(1229).
Meyerhof, G.G. 1976. Bearing capacity and settlement of pile foundations. Jour-
silt and silty clay soil.
nal of Geotechnical Engineering Division, ASCE, 102(GT3): 195–228.
11. The mobilized soil strength parameters underneath the up- Meyerhof, G.G., and Adams, J.I. 1968. The ultimate uplift capacity of foundations.
per helix were higher than those mobilized along the pile Canadian Geotechnical Journal, 5(4): 225–244. doi:10.1139/t68-024.
shaft for the same soil layers. This was attributed to the ex- Mitchell, J.K. 1993. Fundamentals of soil behaviour. John Wiley and Sons Inc.,
tension of the failure surface underneath the upper helix New York, 437p.
beyond the edge of the helices into intact soil that did not Mitsch, M.P., and Clemence, S.P. 1985. The uplift capacity of helix anchors in
sand. In Uplift Behavior of Anchor Foundations in Soil, ASCE Convention
experience any disturbance during pile installation. Conference Proceedings, New York, USA, pp. 26–47.
Mooney, J.M., Adamczak, S.J., and Clemence, S.P. 1985. Uplift capacity of helical
It should be noted that the experimental results are site specific
anchors in clay and silt. In Uplift Behavior of Anchor Foundations in Soil,
and must be used with caution when applied elsewhere. ASCE Convention Conference Proceedings, New York, USA, pp. 48–72.
Narasimha Rao, S., Prasad, Y.V.S.N., Shetty, M.D., and Joshi, V.V. 1989. Uplift
Acknowledgements capacity of screw anchors. Geotechnical Engineering, 20(2): 139–159.
The research described herein received direct support from the Narasimha Rao, S., Prasad, Y.V.S.N., and Shetty, M.D. 1991. The behaviour of
Natural Sciences and Engineering Research Council of Canada model screw piles in cohesive soils. Soils and Foundations, 31: 35–50. doi:10.
3208/sandf1972.31.2_35.
(NSERC), The University of Western Ontario, and Almita Piling Narasimha Rao, S., Prasad, Y.V.S.N., and Veeresh, C. 1993. Behaviour of embed-
Inc., Ponoka, Alberta. The authors would like to thank Erden Ertor
For personal use only.

ded model screw anchors in soft clays. Geotechnique, 43(4): 605–614. doi:10.
(The University of Western Ontario) and David Sparrow and Paul 1680/geot.1993.43.4.605.
Van Der Raadt (AMEC E&I, Inc., Calgary, Alberta) for their support. O’Neill, M.W., and Reese, L.C. 1999. Drilled shafts: construction procedures and
design methods. U.S. Department of Transportation, Federal Highway Ad-
References ministration (FHWA), Report No. FHWA-IF-99-025.
Perko, H. 2001. Energy method for predicting the Installation torque of helical
Adams, J.I., and Hayes, D.C. 1967. The uplift capacity of shallow foundations.
foundations and anchors. In New Technological and Design Developments in
Ontario Hydro Research Quarterly, 19: 1–13.
Deep Foundations, ASCE, Reston, Va., USA, pp. 342–352. doi:10.1061/40511
ASTM. 2007a. Standard specification for welded and seamless steel pipe piles
(288)24.
(A252-98). In Annual Book of ASTM Standards. Vol. 01.01. American Society for
Perko, H.A. 2009. Helical piles: A practical guide to design and installation. 1st
Testing and Materials (ASTM), West Conshohocken, Pa. doi:10.1520/A0252-10.
ASTM. 2007b. Standard test methods for deep foundations under static axial ed. John Wiley & Sons, New Jersey.
compression load (D1143-07). In Annual Book of ASTM Standards. Vol. 04.08. Roy, M., and Lemieux, M. 1986. Long-term behaviour of reconsolidated clay
American Society for Testing and Materials (ASTM), West Conshohocken, Pa. around a driven pile. Canadian Geotechnical Journal, 23(1): 23–29. doi:10.1139/
doi:10.1520/D1143_D1143M. t86-004.
Bishop, A.W. 1966. The strength of soils as engineering materials. Géotechnique, Sakr, M. 2009. Performance of helical piles in oil sand. Canadian Geotechnical
16(2): 91–130. doi:10.1680/geot.1966.16.2.91. Journal, 46(9): 1046–1061. doi:10.1139/T09-044.
Bobbitt, D.E., and Clemence, S.P. 1987. Helical anchors: application and design Sakr, M. 2011. Installation and performance characteristics of high capacity
criteria. In Proceedings of the 9th Southeast Asian Geotechnical Conference, helical piles in cohesionless soils. DFI Journal, 5(1): 39–57. doi:10.1179/dfi.2011.
Bangkok, Thailand, pp. 6-105–6-120. 004.
Bradka, T.D. 1997. Vertical capacity of helical screw anchor piles. Master of Shetsen, I. 1990. Quaternary geology of central Alberta. Alberta Research Coun-
Engineering Report, University of Alberta, Alberta, Canada. cil, Alberta Geological Survey, Map 213.
CFEM. 2006. Canadian Foundation Engineering Manual. 4th ed. Canadian Stark, T.D., and Eid, H.T. 1994. Drained residual strength of cohesive soils. Jour-
Geotechnical Society, BiTech Publisher Ltd., Canada, 488 p. nal of Geotechnical Engineering, 120(5): 856–871. doi:10.1061/(ASCE)0733-
Clemence, S.P., and Pepe, F.D. 1984. Measurement of lateral stress around mul- 9410(1994)120:5(856).
tihelix anchors in sand. Geotechnical Testing Journal, 7(3): 145–152. doi:10. Tappenden, K.M. 2007. Predicting the axial capacity of screw piles installed in
1520/GTJ10490J. western Canadian soils. M.Sc. thesis, University of Alberta, Edmonton, Al-
Davisson, M.T. 1972. High capacity piles. In Proceedings of the Lecture Series, berta, Canada.
Innovations in Foundation Construction, ASCE, Illinois Section, pp. 81–112. Terzaghi, K. 1942. Discussion on progress report of the committee on bearing
Elsherbiny, Z., and El Naggar, M.H. 2013. Axial compressive capacity of helical values of pile foundations, pile driving formulas. In Proceedings of ASCE,
piles from field tests and numerical study. Canadian Geotechnical Journal, Harvard Soil Mechanics Series 17, Vol. 68, pp. 311–323.
50(12): 1191–1203. doi:10.1139/cgj-2012-0487. Terzaghi, K., Peck, R.G., and Mesri, G. 1996. Soil mechanics in engineering prac-
Fellenius, B.H. 1980. The analysis of results from routine pile load tests. Ground tice. John Wiley and Sons Inc., New York, 549p.
Engineering, 13(6): 19–31.
Tsuha, C.deH.C., and Aoki, N. 2010. Relationship between installation torque
Hirany, A., and Kulhawy, F.H. 2002. On the interpretation of drilled foundation and uplift capacity of deep helical piles in sand. Canadian Geotechnical
load test results. In Deep Foundations 2002, An International Perspective on
Journal, 47(6): 635–647. doi:10.1139/T09-128.
Theory, Design, Construction, and Performance, Geotechnical Special Publi-
Vesic, A.S. 1971. Breakout resistance for objects embedded in ocean bottom.
cation No. 116, ASCE, pp. 1018–1028. doi:10.1061/40601(256)71.
Hoyt, R.M., and Clemence, S.P. 1989. Uplift capacity of helical anchors in soil. In Journal of Soil Mechanics and Foundation Division, ASCE, 97(SM9): 1183–
Proceedings of the 12th International Conference on Soil Mechanics and 1205.
Foundation Engineering, Rio de Janeiro, Brazil, Vol. 2, pp. 1019–1022. Vyazmensky, A.M. 2005. Numerical modelling of time dependent pore pressure
Kulhawy, F.H. 1984. Limiting tip and side resistance: fact or fallacy? In Sympo- response induced by helical pile installation. M.E.Sc. thesis, University of
sium on Analysis and Design of Pile Foundations, ASCE, San Francisco, Vol. 1, British Columbia, Vancouver, British Columbia, Canada.
pp. 80–98. Weech, C.N. 2002. Installation and load testing of helical piles in a sensitive
Kulhawy, F.H., and Mayne, P.W. 1990. Manual on estimating soil properties for fine-grained soil. M.E.Sc. thesis, University of British Columbia, Vancouver,
foundation design. Electric Power Research Institute, EPRI, Report No. British Columbia, Canada.
RP1493-6, Palo Alto, Calif., USA, 308 p. Zhang, D.J.Y. 1999. Predicting capacity of helical screw piles in Alberta soils.
Kurian, N.P., and Shah, S.J. 2009. Studies on the behaviour of screw piles by the M.E.Sc. thesis, University of Alberta, Edmonton, Alberta, Canada.

Published by NRC Research Press

View publication stats

You might also like