Download as pdf or txt
Download as pdf or txt
You are on page 1of 5

International Journal of Offshore and Polar Engineering (ISSN 1053-5381) http://www.isope.org/publications/publications.

htm

Downloaded from http://onepetro.org/ijope/article-pdf/25/02/156/2190387/isope-15-25-2-156.pdf?casa_token=D8n8UprjsEIAAAAA:B9V3VvYsoEjahk-dEilQdILaI-xG5tv7uI0FwQe_wGFdFZk4aXxjeDehEr3eKS2OnIL6w5rI by Ohio State University user on 20 June 2024


Copyright © by The International Society of Offshore and Polar Engineers
Vol. 25, No. 2, June 2015, pp. 156–160; http://dx.doi.org/10.17736/ijope.2015.hb14

Comparison of Lateral Behavior of Rock-Socketed Large-Diameter Offshore


Monopiles in Sands with Different Relative Densities
Dongwook Kim
Department of Civil and Environmental Engineering, Incheon National University
Incheon, Korea
Yun Wook Choo*
Department of Civil and Environmental Engineering, Kongju National University
Cheonan-si, Chungnam-do, Korea
Kiseok Kwak
Korea Institute of Civil Engineering and Building Technology
Goyang-si, Gyeonggi-do, Korea

Centrifuge tests were performed to investigate the effect of the relative density of the sandy layer on the lateral response of
6-m-diameter offshore monopiles. The end tips of the monopiles are socketed into rock-bearing layers. For the simulation of the
rock-socketed monopiles in sands with different relative densities (dense and medium dense sand layers), centrifuge tests at an
acceleration of 60 g were conducted using well-instrumented model monopiles under significant lateral loads and bending
moments. Based on the centrifuge test results, the p–y relationship and the initial stiffness (initial gradient of a p–y curve)
changed with an increasing depth for both dense and medium dense sand layers. As a result, the newly developed p–y curves
for rock-socketed large-diameter monopiles in this study were quite different from the existing API (American Petroleum
Institute) p–y curves, which were developed based on relatively small-diameter driven piles. It was found that the initial
stiffness for the medium dense sand layer was significantly lower than that for the dense sand layer at a shallow depth;
however, the ratio of initial stiffness of the medium dense sand layer to that of the dense sand layer decreases significantly as
the depth increases and approaches the stiff rock-bearing layer.

INTRODUCTION because of the limitations of equipment settings in severe offshore


environments. Therefore, centrifuge tests using well-instrumented
Renewable energy produced by wind turbines has now become
monopiles are a good alternative and an effective method of
one of the most promising clean energy resources. At the initial
evaluating monopiles’ behavior.
stage of wind turbine energy production, onshore or inland wind
The p–y method in sands recommended by the American
turbine construction was dominant. However, offshore wind turbines
Petroleum Institute (API) is widely used for the designs of offshore
have been built recently for more steady energy production because
piles, including large-diameter offshore monopiles. The API p–y
the wind quality in terms of wind strength and duration is far better
method (API, 1993) was developed after the work done by Reese
offshore than inland. In addition, the offshore wind turbines are
et al. (1974) based on their experimental results from the lateral
freer from the problems caused by their noise and vibration than
pile load tests of small-diameter piles (a pile diameter of up to
inland wind turbines. The number of offshore construction projects
610 mm) and the simplification process of O’Neill and Murchison
of wind farms keeps increasing worldwide, including off the shores
(1983). Significant diameter effect on p–y formulation has been
of Northern Europe, the United States, China, and Korea. The sizes
experimentally identified in clays (Reese et al., 1975; Stevens and
of generators and blades are increasing for more efficient energy
Audibert, 1979), although only limited literature notes its effect in
production; therefore, larger offshore wind turbine foundations are
sands. However, in the literature (Reese et al., 1975; Stevens and
also required for higher vertical and lateral capacities (resistances)
Audibert, 1979), there was no explanation of the disagreement
against significant vertical and lateral loads and bending moments.
between experimental results and p–y analysis results.
Because the diameters of constructed offshore monopiles are as
Theoretically, Terzaghi and Peck (1948) insisted that the low-
large as 5 m and even larger diameters are required (Achmus et al.,
strain elastic soil stiffness (in this paper, the initial gradient of a
2009), full-scale lateral load tests on the monopiles are practically
p–y curve is defined by the initial soil stiffness ki ) against the
impossible. A full-scale test to estimate the monopile’s vertical and
lateral movement of the pile should be independent of pile diameter.
lateral capacities under offshore conditions is especially difficult
Ashford and Juirnarongrit (2003) also supported the results of
Terzaghi and Peck (1948) based on the full-scale pile load test
*ISOPE Member. results.
Received July 23, 2014; updated and further revised manuscript received by The use of the API p–y method in sands on large-diameter
the editors January 21, 2015. The original version (prior to the final offshore monopile designs without verification of its applicability
updated and revised manuscript) was presented at the Twenty-fourth would lead to unsafe or uneconomical designs. In this study,
International Ocean and Polar Engineering Conference (ISOPE-2014),
Busan, Korea, June 15–20, 2014.
centrifuge tests on well-instrumented rock-socketed large-diameter
KEY WORDS: Offshore wind turbines, offshore foundations, monopole, monopiles in sands were conducted. The lateral behavior of
lateral response, initial stiffness, p–y relationship. monopiles from the centrifuge test results is compared with that
International Journal of Offshore and Polar Engineering, Vol. 25, No. 2, June 2015, pp. 156–160 157

Downloaded from http://onepetro.org/ijope/article-pdf/25/02/156/2190387/isope-15-25-2-156.pdf?casa_token=D8n8UprjsEIAAAAA:B9V3VvYsoEjahk-dEilQdILaI-xG5tv7uI0FwQe_wGFdFZk4aXxjeDehEr3eKS2OnIL6w5rI by Ohio State University user on 20 June 2024


predicted by the API p–y method. In addition, the p–y curves and
their initial stiffnesses, obtained from the experimental results, are
compared with those in the API criteria.

BRIEF SUMMARY OF CENTRIFUGE EXPERIMENT


The model monopiles were selected as follows: the outer diameter
(Dm ) and thickness (tm ) of the model monopile were 101.6 mm
and 2.0 mm in model scale, respectively. Because the g-level of the
centrifuge tests was 60 g, the corresponding outer diameter (Dp )
and thickness (tp ) of the monopile in prototype scale were 6.1 m
and 0.12 m, respectively. The Young’s modulus of the model pile
was 199 GPa, and the moment inertia was 10.06 m4 . The model
monopile is not reused for other tests because of the possibility of
any damages from the previous test.
The sands used in the experiment were nonplastic silica sands
artificially made by crushing quartzite with a hammer crusher. The
main reason to use crushed sands is to minimize particle size scale
effects on the pile lateral behavior. By reducing the particle size,
we expect more realistic interface friction between sands and pile,
which is governed by the ratio of soil particle size and roughness of Fig. 2 Dimensions of the monopile and measurement system in
the pile surface (Choo et al., 2014). The particle size distribution of prototype scale (Dp is the outer diameter of the monopile)
the sands used in the experiment is presented in Fig. 1. The specific
gravity (GS ) was 2.65. The median particle size (D50 5 was 0.22 was 10%) with a moisture content of 10%. An accurately cut real
mm. The curvature and uniformity coefficients (CC and CU 5 were hard rock was located on the bottom of a centrifuge cylindrical
1.11 and 1.96, respectively. The CC and CU values indicate that the container. The completed schematic installation of the monopile
sands were not well-graded soils; however, there was no difficulty is illustrated in Fig. 2. More details of the test preparation and
in forming uniform sand layers having consistent relative density. instrumentation can be found in Choo et al. (2014).
For the formulation of the upper sand layer with different relative
densities (DR ), an automatic sand-raining device was used and we
LATERAL LOAD-DISPLACEMENT RELATIONSHIP
varied the opening size of the hopper and the height of sand falling
from the hopper to the upper layer of sand. The target relative
FROM EXPERIMENT RESULTS AND API
densities of this study were 85% and 55%; however, two relative RECOMMENDATIONS
density levels of 87% and 58% were constructed for the dense and Monopile behavior estimated by the API recommendations is
medium dense upper sand layer. known to be much stiffer than that observed during the numerical
Saturation of the sand layers was conducted after sand layers analyses and experiments (Achmus et al., 2007; Choo et al., 2014).
were formed in dry conditions by letting water flow slowly along However, the much stiffer lateral behavior of piles predicted by the
the tank’s inner face to minimize the disturbance of the sand layers. API method was found for homogeneous sand layers; therefore,
For more saturation, before the main tests, a pilot spinning of the it is interesting to compare the lateral behavior of rock-socketed
centrifuge model (the instrumented soil tank) was performed under
centrifuge acceleration of up to 70 g for more than one hour. After
the pilot spinning, the centrifuge model was ramped down for the
final check-up on the status of the models.
The weathered rock layer was constructed by curing the mixture
of sand and Portland cement (weight percentage of cement to sand

Fig. 3 Lateral load (H) with an increasing lateral displacement


y505 from experimental measurement and API p–y method (H is
the applied lateral load at 33 m above seabed, y505 is the lateral
displacement at 5.5 m above seabed, and DR is the relative density
Fig. 1 Particle size distribution of the sands used in the experiment of a soil layer)
158 Comparison of Lateral Behavior of Rock-Socketed Large-Diameter Offshore Monopiles in Sands with Different Relative Densities

Downloaded from http://onepetro.org/ijope/article-pdf/25/02/156/2190387/isope-15-25-2-156.pdf?casa_token=D8n8UprjsEIAAAAA:B9V3VvYsoEjahk-dEilQdILaI-xG5tv7uI0FwQe_wGFdFZk4aXxjeDehEr3eKS2OnIL6w5rI by Ohio State University user on 20 June 2024


large-diameter monopiles from centrifuge test results to that from In this study, it is assumed that the pu in Eq. 4 is equal to the
the p–y analysis following the API recommendations. In pile pu proposed by the API. The ultimate lateral resistance pu of the
design practices, the p–y analysis method is typically used to check API method in sands is developed from the analytical solution,
the serviceability limit state in terms of pile lateral movement or assuming that the soil at the loading direction side of the pile
rotation. The p–y method originates from a beam–column analysis ultimately exhibits a passive wedge-type failure (Reese et al., 2004).
method of replacing soil with nonlinear p–y springs. The passive wedge-type failure typically occurs when the piles
To obtain a lateral response (load-displacement curve) of the laterally translate without significant rotation. Therefore, the pu of
monopile, lateral load H and lateral displacement y505 were mea- rock-socketed piles in sands could be different to some extent from
sured at 33 m and 5.5 m above the seabed, respectively. The lateral the pu proposed by the API because the deformation mode of the
load was applied far above the seabed to simulate the significant rock-socketed piles involves a combination of lateral translation
lateral load and bending moment induced by wind, waves, and and rotation. This simple assumption of the API about pu would
other environmental loads. The lateral displacements were measured lead to significant disagreement between its pu prediction and the
at multiple locations of the exposed pile; however, because of the true value of pu . Verification of the validity of this assumption,
space limitation, to install displacement transducers, the lowest based on the pu prediction using a more advanced model or on the
possible displacement measurement location was 5.5 m above the pu measurement by sophisticated experiments, is necessary.
seabed in prototype scale. In the experiment of this study, measurement of pu was not
For the two cases (i.e., the different relative densities of the upper possible because of the loading capacity and the limited space of
sand layers), lateral load-displacement curves were obtained and are the cylindrical container of the loading equipment adopted in the
shown in Fig. 3. For a given lateral load, the lateral displacement
(y505 ) was lower in the denser sand layer because the dense sand
layer has a higher stiffness.

METHOD OF OBTAINING EXPERIMENTAL p–y


RELATIONSHIP
Bending moments (M5 along the well-instrumented model
monopile are available from the strain measurements (in pairs at
each vertical level) along the pile’s shaft. To obtain a mathematical
differentiable formulation of the bending moment profile, the
discrete bending moment distribution is fitted to the following form
of a polynomial, as proposed by Wilson (1998):

M = a0 + a1 z + a2 z205 + a3 z3 + a4 z4 + a5 z5 (1)

Additionally, the calculated bending moment profile is used to


calculate the soil reaction force (p5 per length and lateral displace-
ment (y5. The soil reactions are calculated by double differentiation
of the fitted bending moment profile by Eq. 2, and the lateral
displacements are obtained from double integration of the fitting
profile by Eq. 3, with the measured boundary conditions (i.e., three (a)
lateral displacement measurements on the exposed pile).

d2 M
p=− (2)
dz2
Z Z  Z Z M 
y=  dz dz = dz dz (3)
EI

where  is the curvature of the monopile (the ratio of the difference


between compression and tension strains measured at a depth of
interest to the distance between these two strain gages), M is
the bending moment at a depth of interest z, EI is the flexural
modulus of the monopile, p is the soil reaction per unit length of
the monopile, and y is the lateral displacement of the pile at a
depth of interest.

INITIAL STIFFNESS PROFILE FROM MEASUREMENT


The initial soil stiffness ki at a certain depth is obtained from
regression analysis using experimental results for the following given
hyperbolic equation with the assumed ultimate lateral resistance pu : (b)
y Fig. 4 Best-fit p–y curves based on the experimental results for
p= (4)
1 y different embedded depths: (a) dense sand upper layer and (b)
+
ki p u medium dense sand upper layer
International Journal of Offshore and Polar Engineering, Vol. 25, No. 2, June 2015, pp. 156–160 159

Downloaded from http://onepetro.org/ijope/article-pdf/25/02/156/2190387/isope-15-25-2-156.pdf?casa_token=D8n8UprjsEIAAAAA:B9V3VvYsoEjahk-dEilQdILaI-xG5tv7uI0FwQe_wGFdFZk4aXxjeDehEr3eKS2OnIL6w5rI by Ohio State University user on 20 June 2024


centrifuge. Although the pu that was used may be different from
the real value, the experimental p–y curves developed in this paper
match well the experimental results within the measured range of
lateral displacements.
Unfortunately, strain measurements at all levels were not pos-
sible. The strain measurements within rock layers were erratic
or unobtainable because the strain levels of the rock-socketed
part were very small and negligible. However, most of the strain
measurements within the sand layers were successful. The fitted
equations (Eq. 4) and the p–y relationship from the experimental
measurements are plotted in Fig. 4 for dense and medium dense
sand layers, respectively.
As shown in Fig. 4, the experimental data fit well to the
hyperbolic p–y relationship curves. The initial soil stiffness ki
increases with depth for both soil profiles; yet, the increasing trends
of the initial soil stiffness of the two soil profiles are different. The
variation in the initial soil stiffness for both cases is plotted in Fig. 6 Lateral load H with an increasing lateral displacement y505
Fig. 5. from the experimental measurement and p–y method with revised
The initial soil stiffness ki varies with the friction angle (or initial stiffness for dense and medium dense sand upper layers
relative density) of the soil and water table conditions (i.e., whether
a soil layer is placed below or above the water table). The API
recommendations (API, 1993) propose a proportional relationship cases becomes smaller as the depth z becomes closer to the rock
between the initial soil stiffness and depth (Fig. 5). The initial soil layers. The effect of a strong, stiff layer on the p–y relationship
stiffness ki profiles of the dense and medium dense sand layers do (including ki 5 of the pile and the adjacent relatively weak layer is
not match those recommended by the API. The ki values from also mentioned by Yang and Jeremic (2005), who revealed that
the API recommendations are much higher than those obtained the ki of multilayers is interactively influenced by neighboring
from the experiment. Davisson (1963) also mentioned that the layers (as we found in this study). “Layers with different stiffness
initial soil stiffness increases linearly with depth for granular soils, effects on p–y curves” were initially discussed by Georgiadis
normally loaded silts, and clays. On the other hand, Yan and (1983) and modified by Yang and Jeremic (2005) based on the
Byrne (1992) claimed that the ki increases with increasing depth, rigorous three-dimensional finite element analysis results. Yang and
but the relationship between ki and depth is not linear. From the Jeremic (2005) concluded that the p–y relationship of a pile within
comparison between soil elastic moduli and depths, they found that a soft layer over a stiff layer becomes stiffer than the curve within
the increment of ki with an increasing depth tends to decrease. a single homogeneous soft layer.
In addition, for the two cases (the dense and medium dense sand Another issue of ki is the existence of diameter effect on ki . For
upper layers), as the depth z becomes deeper and closer to relatively a given depth, Carter (1984) and Ling (1988) found that the initial
stiff rock layers, the gradient ki of p with respect to y becomes soil stiffness ki increases with an increasing pile diameter. The
much stiffer than for that installed in a single sand layer without results were derived from the literature review for pile diameter
stiff rock layers. The experimental results without rock layers are ranging from approximately 8 cm to 90 cm. However, the centrifuge
available in Choo et al. (2014). Furthermore, the difference in test results derived in the present study show a disagreement with
ki between the dense sand layer and medium dense sand layer the results by Carter (1984) and Ling (1988). A further study
should be performed regarding the pile diameter effect on the p–y
relationship.
Based on the obtained p–y relationship (Eq. 4) from the experi-
ment, a lateral load-displacement relationship is examined using
the p–y numerical analysis. The results are plotted in Fig. 6. The
p–y analysis using the experimental p–y relationship tends to
slightly overestimate the lateral displacement y505 within a small
magnitude of lateral load H (approximately H < 14 MN) and tends
to underestimate the lateral displacement for H > 14 MN. However,
the experimental p–y relationships reasonably capture the overall
lateral behavior of the large-diameter monopiles.

CONCLUSIONS
The applicability of the API p–y method in sands on rock-
socketed large-diameter offshore monopiles is examined based on
centrifuge test results. It was found that the API p–y method rea-
sonably captures the apparent lateral load-displacement relationship
from centrifuge test results within a small lateral displacement range
(lateral displacement within 0.05 m for a monopile with a 6.1 m
diameter). However, for larger lateral displacements, the API p–y
method results in a much stiffer lateral behavior compared to cen-
Fig. 5 Variation in initial soil stiffness with an increasing depth trifuge test results. In addition, a comparison of initial soil stiffness
from the seabed for dense and medium dense sand upper layers in a p–y curve is made between experimental results and the API
160 Comparison of Lateral Behavior of Rock-Socketed Large-Diameter Offshore Monopiles in Sands with Different Relative Densities

Downloaded from http://onepetro.org/ijope/article-pdf/25/02/156/2190387/isope-15-25-2-156.pdf?casa_token=D8n8UprjsEIAAAAA:B9V3VvYsoEjahk-dEilQdILaI-xG5tv7uI0FwQe_wGFdFZk4aXxjeDehEr3eKS2OnIL6w5rI by Ohio State University user on 20 June 2024


recommendations. The initial soil stiffness ki of p–y curves from Tests," Geotech Test J, ASTM, 37(1), 107–120.
the centrifuge tests was much lower than that proposed by the API. http://dx.doi.org/10.1520/GTJ20130081.
For the piles installed along multiple layers with significantly dif- Davisson, MT (1963). “Estimating Buckling Loads for Piles,” Proc
ferent soil or rock stiffnesses, the initial soil stiffness in p–y curves 2nd Pan Am Conf Soil Mech Found Eng, Sociedad Mexicana de
of the softer layer increases significantly as the depth approaches Mecánica de Suelos, Mexico City, Mexico, 1, 351–371.
the much stiffer layer. The ki in the sand layer with increasing Georgiadis, M (1983). “Development of p-y Curves for Layered
depth for rock-socketed piles is significantly different from that Soils,” Proc Geotech Pract Offshore Eng, Austin, TX, USA,
for piles embedded in uniform sand layers. Previous researchers ASCE, 536–545.
claimed that the ki for a uniform sandy layer either increases
Ling, LF (1988). Back Analysis of Lateral Load Tests on Piles, MS
linearly or increases with reduced increment with increasing depth.
Thesis, Civil Engineering Department, University of Auckland,
Thus, more research is required to quantify the effect of difference
New Zealand0
in stiffness between adjacent soil or rock layers on the initial soil
stiffness of the p–y relationship. O’Neill, MW, and Murchison, JM (1983). An Evaluation of p–y
Relationship in Sands, A Report to the American Petroleum
Institute, PRAC 82-41-1. Also Department of Civil Engineer-
ACKNOWLEDGEMENTS ing Research Report No GT-DF02-83, University of Houston,
The study presented in this paper was supported by Grant University Park, TX, USA.
R&D/10 Technology Innovation E04 from the Offshore Wind- Reese, LC, Cox, WR, and Koop, FD (1974). “Analysis of Laterally
Energy Foundation System Project funded by the Ministry of Land, Loaded Piles in Sand,” Proc 6th Annu Offshore Technol Conf,
Infrastructure and Transport (MOLIT). The authors acknowledge Houston, TX, USA, OTC 2080, 473–483.
all the support from the MOLIT. http://dx.doi.org/10.4043/2080-MS.
Reese, LC, Cox, WR, and Koop, FD (1975). “Field Testing and
REFERENCES Analysis of Laterally Loaded Piles in Stiff Clay,” Proc 7th Annu
Offshore Technol Conf, Houston, TX, USA, OTC 2312, 671–690.
Achmus, M, Abdel-Rahman, K, and Kuo, Y-S (2007). “Behavior of http://dx.doi.org/10.4043/2312-MS.
Large Diameter Monopiles Under Cyclic Horizontal Loading,” Reese, LC, Wang, ST, Isenhower, WM, and Arrellaga, JA (2004).
In Proc 12th Int Colloquium Struct Geotech Eng, Cairo, Egypt, A Program for the Analysis of Piles and Drilled Shafts Under
11 pp. Lateral Loads, Ensoft, Inc, Austin, TX, USA.
Achmus, M, Kuo, YS, and Abdel-Rahman, K (2009). “Behavior of Stevens, JB, and Audibert, JME (1979). “Re-examination of p–y
Monopile Foundations Under Cyclic Lateral Load,” Comput Curve Formulations,” Proc 11th Annu Offshore Technol Conf,
Geotech, 36(5), 725–735. Houston, TX, USA, OTC 3402, 397–403.
http://dx.doi.org/10.1016/j.compgeo.2008.12.003. http://dx.doi.org/10.4043/3402-MS.
API (1993). Recommended Practice for Planning, Designing and Terzaghi, K, and Peck, RB (1948). Soil Mechanics in Engineering
Constructing Fixed Offshore Platforms: Working Stress Design, Practice, 1st Ed, John Wiley and Sons, 566 pp.
API-RP-2A-WSD, 20th ed, American Petroleum Institute. Wilson, D (1998). Soil-Pile-Superstructure Interaction in Liquefying
Ashford, S, and Juirnarongrit, T (2003). “Evaluation of Pile Diame- Sand and Soft Clay, PhD Dissertation, University of California
ter Effect on Initial Modulus of Sugrade Reaction,” J Geotech at Davis, CA, USA.
Geoenviron Eng, ASCE, 129(3), 234–242. Yan, L, and Byrne, PM (1992). “Lateral Pile Response to Mono-
http://dx.doi.org/10.1061/(ASCE)1090-0241(2003)129:3(234). tonic Pile Head Loading,” Can Geotech J, 29(6), 955–970.
Carter, DP (1984). A Nonlinear Soil Model for Predicting Lateral http://dx.doi.org/10.1139/t92-106.
Pile Response, Technical Report No 359, Civil Engineering Yang, Z, and Jeremic, B (2005). “Study of Soil Layering Effects
Department, University of Auckland, New Zealand. on Lateral Loading Behavior of Piles,” J Geotech Geoenviron
Choo, YW, et al. (2014). “Lateral Response of Large-Diameter Eng, ASCE, 131(6), 762–770.
Monopiles for Offshore Wind Turbines from Centrifuge Model http://dx.doi.org/10.1061/(ASCE)1090-0241(2005)131:6(762).

You might also like